21.03.2015 Views

Introduction to Fungi, Third Edition

Introduction to Fungi, Third Edition

Introduction to Fungi, Third Edition

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

This page intentionally left blank


<strong>Introduction</strong><strong>to</strong> <strong>Fungi</strong><br />

This new edition of the universally acclaimed and<br />

widely used textbook on fungal biology has been<br />

completely rewritten, drawing directly on the<br />

authors’ research and teaching experience. The<br />

text takes account of the rapid and exciting<br />

progress that has been made in the taxonomy, cell<br />

and molecular biology, biochemistry, pathology<br />

and ecology of the fungi. Features of taxonomic<br />

significance are integrated with natural functions,<br />

including their relevance <strong>to</strong> human affairs. Special<br />

emphasis is placed on the biology and control of<br />

human and plant pathogens, providing a vital<br />

link between fundamental and applied mycology.<br />

The book is richly illustrated throughout with<br />

specially prepared drawings and pho<strong>to</strong>graphs,<br />

based on living material. Illustrated life cycles are<br />

provided, and technical terms are clearly explained.<br />

Extensive reference is made <strong>to</strong> recent literature and<br />

developments, and the emphasis throughout is on<br />

whole-organism biology from an integrated, multidisciplinary<br />

perspective.<br />

John Webster is Professor Emeritus of the School of<br />

Biosciences at the University of Exeter, UK.<br />

Roland W.S. Weber was a Lecturer in the Department<br />

of Biotechnology at the University of Kaiserslautern,<br />

Germany, and is currently at the Fruit Experiment<br />

station (OVB) in Jork, Germany.


<strong>Introduction</strong> <strong>to</strong> <strong>Fungi</strong><br />

JohnWebster<br />

University of Exeter<br />

and<br />

Roland Weber<br />

University of Kaiserslautern<br />

<strong>Third</strong> <strong>Edition</strong>


CAMBRIDGE UNIVERSITY PRESS<br />

Cambridge, New York, Melbourne, Madrid, Cape Town, Singapore, São Paulo<br />

Cambridge University Press<br />

The Edinburgh Building, Cambridge CB2 8RU, UK<br />

Published in the United States of America by Cambridge University Press, New York<br />

www.cambridge.org<br />

Information on this title: www.cambridge.org/9780521807395<br />

© J. Webster and R. W. S. Weber 2007<br />

This publication is in copyright. Subject <strong>to</strong> statu<strong>to</strong>ry exception and <strong>to</strong> the provision of<br />

relevant collective licensing agreements, no reproduction of any part may take place<br />

without the written permission of Cambridge University Press.<br />

First published in print format<br />

2007<br />

ISBN-13 978-0-511-27783-2 eBook (EBL)<br />

ISBN-10 0-511-27783-0 eBook (EBL)<br />

ISBN-13 978-0-521-80739-5 hardback<br />

ISBN-10 0-521-80739-5 hardback<br />

ISBN-13 978-0-521-01483-0 paperback<br />

ISBN-10 0-521-01483-2 paperback<br />

Cambridge University Press has no responsibility for the persistence or accuracy of urls<br />

for external or third-party internet websites referred <strong>to</strong> in this publication, and does not<br />

guarantee that any content on such websites is, or will remain, accurate or appropriate.


To Philip M. Booth


Contents<br />

Preface <strong>to</strong> the first edition<br />

Preface <strong>to</strong> the second edition<br />

Preface <strong>to</strong> the third edition<br />

Acknowledgements<br />

page xiii<br />

xv<br />

xvii<br />

xix<br />

Chapter 1 <strong>Introduction</strong> 1<br />

1.1 What are fungi? 1<br />

1.2 Physiology of the growing hypha 3<br />

1.3 Hyphal aggregates 14<br />

1.4 Spores of fungi 22<br />

1.5 Taxonomy of fungi 32<br />

Chapter 2 Pro<strong>to</strong>zoa: Myxomycota (slime moulds) 40<br />

2.1 <strong>Introduction</strong> 40<br />

2.2 Acrasiomycetes: acrasid cellular slime moulds 40<br />

2.3 Dictyosteliomycetes: dictyostelid slime moulds 41<br />

2.4 Pro<strong>to</strong>steliomycetes: pro<strong>to</strong>stelid plasmodial<br />

slime moulds 45<br />

2.5 Myxomycetes: true (plasmodial) slime moulds 47<br />

Chapter 3 Pro<strong>to</strong>zoa: Plasmodiophoromycota 54<br />

3.1 <strong>Introduction</strong> 54<br />

3.2 Plasmodiophorales 54<br />

3.3 Control of diseases caused by Plasmodiophorales 62<br />

3.4 Hap<strong>to</strong>glossa (Hap<strong>to</strong>glossales) 64<br />

Chapter 4 Straminipila: minor fungal phyla 67<br />

4.1 <strong>Introduction</strong> 67<br />

4.2 The straminipilous flagellum 68<br />

4.3 Hyphochytriomycota 70<br />

4.4 Labyrinthulomycota 71<br />

Chapter 5 Straminipila: Oomycota 75<br />

5.1 <strong>Introduction</strong> 75<br />

5.2 Saprolegniales 79<br />

5.3 Pythiales 95<br />

5.4 Peronosporales 115<br />

5.5 Sclerosporaceae 125


viii<br />

CONTENTS<br />

Chapter 6 Chytridiomycota 127<br />

6.1 <strong>Introduction</strong> 127<br />

6.2 Chytridiales 134<br />

6.3 Spizellomycetales 145<br />

6.4 Neocallimastigales (rumen fungi) 150<br />

6.5 Blas<strong>to</strong>cladiales 153<br />

6.6 Monoblepharidales 162<br />

Chapter 7 Zygomycota 165<br />

7.1 <strong>Introduction</strong> 165<br />

7.2 Zygomycetes: Mucorales 165<br />

7.3 Examples of Mucorales 180<br />

7.4 Zoopagales 200<br />

7.5 En<strong>to</strong>mophthorales 202<br />

7.6 Glomales 217<br />

7.7 Trichomycetes 222<br />

Chapter 8 Ascomycota (ascomycetes) 226<br />

8.1 <strong>Introduction</strong> 226<br />

8.2 Vegetative structures 226<br />

8.3 Life cycles of ascomycetes 228<br />

8.4 Conidia of ascomycetes 230<br />

8.5 Conidium production in ascomycetes 231<br />

8.6 Development of asci 236<br />

8.7 Types of fruit body 245<br />

8.8 Fossil ascomycetes 246<br />

8.9 Scientific and economic significance<br />

of ascomycetes 246<br />

8.10 Classification 247<br />

Chapter 9 Archiascomycetes 250<br />

9.1 <strong>Introduction</strong> 250<br />

9.2 Taphrinales 251<br />

9.3 Schizosaccharomycetales 253<br />

9.4 Pneumocystis 259<br />

Chapter 10 Hemiascomycetes 261<br />

10.1 <strong>Introduction</strong> 261<br />

10.2 Saccharomyces (Saccharomycetaceae) 263<br />

10.3 Candida (anamorphic Saccharomycetales) 276<br />

10.4 Pichia (Saccharomycetaceae) 281<br />

10.5 Galac<strong>to</strong>myces (Dipodascaceae) 281<br />

10.6 Saccharomycopsis (Saccharomycopsidaceae) 282<br />

10.7 Eremothecium (Eremotheciaceae) 284


CONTENTS<br />

ix<br />

Chapter 11 Plec<strong>to</strong>mycetes 285<br />

11.1 <strong>Introduction</strong> 285<br />

11.2 Ascosphaerales 286<br />

11.3 Onygenales 289<br />

11.4 Eurotiales 297<br />

Chapter 12 Hymenoascomycetes: Pyrenomycetes 315<br />

12.1 <strong>Introduction</strong> 315<br />

12.2 Sordariales 315<br />

12.3 Xylariales 332<br />

12.4 Hypocreales 337<br />

12.5 Clavicipitales 348<br />

12.6 Ophios<strong>to</strong>matales 364<br />

12.7 Microascales 368<br />

12.8 Diaporthales 373<br />

12.9 Magnaporthaceae 377<br />

12.10 Glomerellaceae 386<br />

Chapter 13 Hymenoascomycetes: Erysiphales 390<br />

13.1 <strong>Introduction</strong> 390<br />

13.2 Phylogenetic aspects 392<br />

13.3 Blumeria graminis 393<br />

13.4 Erysiphe 401<br />

13.5 Podosphaera and Sphaerotheca 404<br />

13.6 Sawadaea 405<br />

13.7 Phyllactinia and Leveillula 405<br />

13.8 Control of powdery mildew diseases 408<br />

Chapter 14<br />

Hymenoascomycetes: Pezizales<br />

(operculate discomycetes) 414<br />

14.1 <strong>Introduction</strong> 414<br />

14.2 Pyronema (Pyronemataceae) 415<br />

14.3 Aleuria (Pyronemataceae) 417<br />

14.4 Peziza (Pezizaceae) 419<br />

14.5 Ascobolus (Ascobolaceae) 419<br />

14.6 Helvella (Helvellaceae) 423<br />

14.7 Tuber (Tuberaceae) 423<br />

14.8 Morchella (Morchellaceae) 427<br />

Chapter 15<br />

Hymenoascomycetes: Helotiales<br />

(inoperculate discomycetes) 429<br />

15.1 <strong>Introduction</strong> 429<br />

15.2 Sclerotiniaceae 429<br />

15.3 Dermateaceae 439


x<br />

CONTENTS<br />

15.4 Rhytismataceae 440<br />

15.5 Other representatives of the Helotiales 442<br />

Chapter 16<br />

Lichenized fungi (chiefly<br />

Hymenoascomycetes: Lecanorales) 446<br />

16.1 <strong>Introduction</strong> 446<br />

16.2 General aspects of lichen biology 447<br />

16.3 Lecanorales 455<br />

Chapter 17 Loculoascomycetes 459<br />

17.1 <strong>Introduction</strong> 459<br />

17.2 Pleosporales 460<br />

17.3 Dothideales 480<br />

Chapter 18 Basidiomycota 487<br />

18.1 <strong>Introduction</strong> 487<br />

18.2 Basidium morphology 487<br />

18.3 Development of basidia 488<br />

18.4 Basidiospore development 490<br />

18.5 The mechanism of basidiospore discharge 493<br />

18.6 Numbers of basidiospores 495<br />

18.7 Basidiospore germination and hyphal growth 496<br />

18.8 Asexual reproduction 501<br />

18.9 Mating systems in basidiomycetes 506<br />

18.10 Fungal individualism: vegetative incompatibility<br />

between dikaryons 510<br />

18.11 Relationships 511<br />

18.12 Classification 512<br />

Chapter 19 Homobasidiomycetes 514<br />

19.1 <strong>Introduction</strong> 514<br />

19.2 Structure and morphogenesis of basidiocarps 517<br />

19.3 Importance of homobasidiomycetes 525<br />

19.4 Euagarics clade 532<br />

19.5 Bole<strong>to</strong>id clade 555<br />

19.6 Polyporoid clade 560<br />

19.7 Russuloid clade 566<br />

19.8 Thelephoroid clade 572<br />

19.9 Hymenochae<strong>to</strong>id clade 573<br />

19.10 Cantharelloid clade 574<br />

19.11 Gomphoid phalloid clade 575<br />

Chapter 20 Homobasidiomycetes: gasteromycetes 577<br />

20.1 <strong>Introduction</strong> 577<br />

20.2 Evolution and phylogeny of gasteromycetes 578


CONTENTS<br />

xi<br />

20.3 Gasteromycetes in the euagarics clade 581<br />

20.4 Gasteromycetes in the bole<strong>to</strong>id clade 585<br />

20.5 Gasteromycetes in the gomphoid phalloid clade 588<br />

Chapter 21 Heterobasidiomycetes 593<br />

21.1 <strong>Introduction</strong> 593<br />

21.2 Cera<strong>to</strong>basidiales 594<br />

21.3 Dacrymycetales 598<br />

21.4 Auriculariales 601<br />

21.5 Tremellales 604<br />

Chapter 22 Urediniomycetes: Uredinales (rust fungi) 609<br />

22.1 Urediniomycetes 609<br />

22.2 Uredinales: the rust fungi 609<br />

22.3 Puccinia graminis, the cause of black stem rust 620<br />

22.4 Other cereal rusts 627<br />

22.5 Puccinia and Uromyces 629<br />

22.6 Other members of the Pucciniaceae 631<br />

22.7 Melampsoraceae 634<br />

Chapter 23<br />

Ustilaginomycetes: smut fungi and<br />

their allies 636<br />

23.1 Ustilaginomycetes 636<br />

23.2 The ‘true’ smut fungi (Ustilaginomycetes) 636<br />

23.3 Microbotryales (Urediniomycetes) 652<br />

23.4 Exobasidiales (Ustilaginomycetes) 655<br />

Chapter 24 Basidiomycete yeasts 658<br />

24.1 <strong>Introduction</strong> 658<br />

24.2 Heterobasidiomycete yeasts 660<br />

24.3 Urediniomycete yeasts 666<br />

24.4 Ustilaginomycete yeasts 670<br />

Chapter 25<br />

Anamorphic fungi (nema<strong>to</strong>phagous and<br />

aquatic forms) 673<br />

25.1 Nema<strong>to</strong>phagous fungi 673<br />

25.2 Aquatic hyphomycetes (Ingoldian fungi) 685<br />

25.3 Aero-aquatic fungi 696<br />

References 702<br />

Index 817<br />

Colour plate section appears between pages 412 and 413


Preface <strong>to</strong> the first edition<br />

There are several available good textbooks of<br />

mycology, and some justification is needed for<br />

publishing another. I have long been convinced<br />

that the best way <strong>to</strong> teach mycology, and indeed<br />

all biology, is <strong>to</strong> make use, wherever possible, of<br />

living material. Fortunately with fungi, provided<br />

one chooses the right time of the year, a wealth<br />

of material is readily available. Also by use<br />

of cultures and by infecting material of plant<br />

pathogens in the glasshouse or by maintaining<br />

pathological plots in the garden, it is possible <strong>to</strong><br />

produce material at almost any time. I have<br />

therefore tried <strong>to</strong> write an introduction <strong>to</strong> fungi<br />

which are easily available in the living state, and<br />

have tried <strong>to</strong> give some indication of where they<br />

can be obtained. In this way I hope <strong>to</strong> encourage<br />

students <strong>to</strong> go in<strong>to</strong> the field and look for fungi<br />

themselves. The best way <strong>to</strong> begin is <strong>to</strong> go with<br />

an expert, or <strong>to</strong> attend a Fungus Foray such<br />

as those organized in the spring and autumn<br />

by mycological and biological societies. I owe<br />

much of my own mycological education <strong>to</strong> such<br />

friendly gatherings. A second aim has been <strong>to</strong><br />

produce original illustrations of the kind that a<br />

student could make for himself from simple<br />

preparations of living material, and <strong>to</strong> illustrate<br />

things which he can verify for himself. For this<br />

reason I have chosen not <strong>to</strong> use electron micrographs,<br />

but <strong>to</strong> make drawings based on them.<br />

The problem of what <strong>to</strong> include has been<br />

decided on the criterion of ready availability.<br />

Where an uncommon fungus has been included<br />

this is because it has been used <strong>to</strong> establish some<br />

important fact or principle. A criticism which I<br />

must accept is that no attempt has been made <strong>to</strong><br />

deal with <strong>Fungi</strong> Imperfecti as a group. This is not<br />

because they are not common or important but<br />

that <strong>to</strong> have included them would have made the<br />

book much longer. To mitigate this shortcoming<br />

I have described the conidial states of some<br />

Ascomycotina rather fully, <strong>to</strong> include reference<br />

<strong>to</strong> some of the form-genera which have been<br />

linked with them. A more difficult problem has<br />

been <strong>to</strong> know which system of classification <strong>to</strong><br />

adopt. I have finally chosen the ‘General Purpose<br />

Classification’ proposed by Ainsworth, which is<br />

adequate for the purpose of providing a framework<br />

of reference. I recognize that some might<br />

wish <strong>to</strong> classify fungi differently, but see no great<br />

merit in burdening the student with the arguments<br />

in favour of this or that system.<br />

Because the evidence for the evolutionary<br />

origins of fungi is so meagre I have made only<br />

scant reference <strong>to</strong> the speculations which have<br />

been made on this <strong>to</strong>pic. There are so many<br />

observations which can be verified, and for this<br />

reason I have preferred <strong>to</strong> leave aside those<br />

which never will.<br />

The literature on fungi is enormous, and<br />

expanding rapidly. Many undergraduates do not<br />

have much time <strong>to</strong> check original publications.<br />

However, since the book is intended as an<br />

introduction I have tried <strong>to</strong> give references <strong>to</strong><br />

some of the more recent literature, and at the<br />

same time <strong>to</strong> quote the origins of some of the<br />

statements made.<br />

Exeter, 27 April 1970<br />

J.W.


Preface <strong>to</strong> the second edition<br />

In revising the first edition, which was first<br />

published about ten years ago, I have taken the<br />

opportunity <strong>to</strong> give a more complete account<br />

of the Myxomycota, and <strong>to</strong> give a more general<br />

introduction <strong>to</strong> the Eumycota. An account<br />

has also been given of some conidial fungi,<br />

as exemplified by aquatic <strong>Fungi</strong> Imperfecti,<br />

nema<strong>to</strong>phagous fungi and seed-borne fungi.<br />

The taxonomic framework has been based on<br />

Volumes IVA and IVB of Ainsworth, Sparrow<br />

and Sussman’s The <strong>Fungi</strong>: An Advanced Treatise<br />

(Academic Press, 1973).<br />

Exeter, January 1979<br />

J.W.


Preface <strong>to</strong> the third edition<br />

Major advances, especially in DNA-based technology,<br />

have catalysed a sheer explosion of mycological<br />

knowledge since the second edition of<br />

<strong>Introduction</strong> <strong>to</strong> <strong>Fungi</strong> was published some<br />

25 years ago. As judged by numbers of publications,<br />

the field of molecular phylogeny, i.e. the<br />

computer-aided comparison of homologous DNA<br />

or protein sequences, must be at the epicentre of<br />

these developments. As a result, information is<br />

now available <strong>to</strong> facilitate the establishment of<br />

taxonomic relationships between organisms or<br />

groups of organisms on a firmer basis than that<br />

previously assumed from morphological resemblance.<br />

This has in turn led <strong>to</strong> revised systems of<br />

classification and provided evidence on which <strong>to</strong><br />

base opinions on the possible evolutionary origin<br />

of fungal groups. We have attempted <strong>to</strong> reflect<br />

some of these advances in this edition. In general<br />

we have followed the outline system of classification<br />

set out in The Mycota Volume VII (Springer-<br />

Verlag) and the Dictionary of the <strong>Fungi</strong> (ninth<br />

edition, CABI Publishing). However, the main<br />

emphasis of our book remains that of presenting<br />

the fungi in a sensible biological context which<br />

can be unders<strong>to</strong>od by students, and therefore<br />

some fungi have been treated along with<br />

taxonomically separate groups if these share<br />

fundamental biological principles. Examples<br />

include Microbotryum, which is treated <strong>to</strong>gether<br />

with smut fungi rather than the rusts <strong>to</strong> which<br />

it belongs taxonomically, or Hap<strong>to</strong>glossa, which<br />

we discuss alongside Plasmodiophora rather than<br />

with the Oomycota.<br />

Molecular phylogeny has been instrumental<br />

in clarifying the relationships of anamorphic<br />

fungi (fungi imperfecti), presenting an opportunity<br />

<strong>to</strong> integrate their treatment with sexually<br />

reproducing relatives. There are only a few<br />

groups such as nema<strong>to</strong>phagous fungi and the<br />

aquatic and aero-aquatic hyphomycetes which<br />

we continue <strong>to</strong> treat as ecological entities rather<br />

than scattered among ascomycetes and basidiomycetes.<br />

Similarly, the gasteromycetes, clearly<br />

an unnatural assemblage, are described <strong>to</strong>gether<br />

because of their unifying biological features.<br />

However, in all these cases taxonomic affinities<br />

are indicated where known. We have also<br />

included several groups now placed well outside<br />

the <strong>Fungi</strong>, such as the Oomycota (Straminipila)<br />

and Myxomycota and Plasmodiophoromycota<br />

(Pro<strong>to</strong>zoa). This is because of their biological<br />

and economic importance and because they have<br />

been and continue <strong>to</strong> be studied by mycologists.<br />

There have been major advances in other<br />

areas of research, notably the molecular cell<br />

biology of the two yeasts Saccharomyces and<br />

Schizosaccharomyces, ‘model organisms’ which<br />

have a bearing far beyond mycology. Further,<br />

much exciting progress is being made in elucidating<br />

the molecular aspects of the infection<br />

biology of human and plant pathogens, and in<br />

developing fungi for biotechnology. These trends<br />

are represented in the current edition.<br />

Nevertheless, the fundamental concept of<br />

<strong>Introduction</strong> <strong>to</strong> <strong>Fungi</strong> remains that of the previous<br />

two editions: <strong>to</strong> place an organism in its<br />

taxonomic context while discussing as many<br />

relevant aspects of its biology as possible in a<br />

holistic manner. Many of the illustrations are<br />

based on original line drawings because we<br />

believe that these can readily portray an understanding<br />

of structure and that drawing as a<br />

record of interpretation is a good discipline.<br />

However, we have also extended the use of<br />

pho<strong>to</strong>graphs, and we now provide illustrated<br />

life cycles because these are more easily unders<strong>to</strong>od.<br />

As before, our choice of illustrated species<br />

has been influenced by the ready availability of<br />

material, enabling students and their teachers<br />

<strong>to</strong> examine living fungi, which is a corners<strong>to</strong>ne<br />

of good teaching. At their first introduction most<br />

technical terms have been printed in bold, their<br />

meanings explained and their derivations given.<br />

The page numbers where these definitions are<br />

given have been highlighted in the index.<br />

The discipline of mycology has evolved and<br />

diversified so enormously in recent decades that<br />

it is now a daunting task for individual authors<br />

<strong>to</strong> give a balanced, integrated account of the<br />

fungi. Of course, there will be omissions or


xviii<br />

PREFACE TO THE THIRD EDITION<br />

misrepresentations in a work of this scale, and<br />

we offer our apologies <strong>to</strong> those who feel that<br />

their work or that of others has not been<br />

adequately covered. At the same time, it has<br />

been a fascinating experience for us <strong>to</strong> write this<br />

book, and we have thoroughly enjoyed the<br />

immense diversity of approaches and ideas<br />

which make mycology such a vibrant discipline<br />

at present. We hope <strong>to</strong> have conveyed some of its<br />

fascination <strong>to</strong> the reader in the text and by<br />

referring <strong>to</strong> as many original publications as<br />

possible.<br />

Exeter and Kaiserslautern, 1 March 2006<br />

J.W. and R.W.S.W.


Acknowledgements<br />

We are indebted <strong>to</strong> many people who have<br />

helped us in our extensive revisions <strong>to</strong> <strong>Introduction</strong><br />

<strong>to</strong> <strong>Fungi</strong>. This edition is dedicated <strong>to</strong> Mr Philip M.<br />

Booth in profound gratitude for his financial<br />

support and his encouragement over many years.<br />

We have acknowledged in the figure legends<br />

the many friends and colleagues who have<br />

responded so enthusiastically <strong>to</strong> our call for<br />

help by providing us with illustrations, sometimes<br />

previously unpublished, and we thank<br />

numerous publishing houses for permission <strong>to</strong><br />

include published figures. We thank Caroline<br />

Huxtable and Rob Ford (Exeter University<br />

Library) and Jennifer Mergel and Petra Tremmel<br />

(Kaiserslautern University Library) for help<br />

beyond the call of duty in obtaining interlibrary<br />

loans. Dr Wolf-Rüdiger Arendholz and<br />

Dr Roger T.A. Cook have read the entire manuscript<br />

or parts of it, and their feedback and<br />

corrections have been most valuable <strong>to</strong> us. We are<br />

immensely grateful <strong>to</strong> Professors Heidrun and<br />

Timm Anke (Kaiserslautern) for their support of<br />

this project, their encouragement and for providing<br />

such a stimulating environment for research<br />

and teaching of fungal biology.<br />

By far the heaviest <strong>to</strong>ll has been paid by our<br />

families and friends who have had only cursory<br />

sightings of us during the past six years. We owe<br />

a debt of gratitude <strong>to</strong> them for their patient<br />

forbearance and unwavering support.


1<br />

<strong>Introduction</strong><br />

1.1 What are fungi?<br />

About 80 000 <strong>to</strong> 120 000 species of fungi have been<br />

described <strong>to</strong> date, although the <strong>to</strong>tal number of<br />

species is estimated at around 1.5 million<br />

(Hawksworth, 2001; Kirk et al., 2001). This would<br />

render fungi one of the least-explored biodiversity<br />

resources of our planet. It is no<strong>to</strong>riously difficult<br />

<strong>to</strong> delimit fungi as a group against other eukaryotes,<br />

and debates over the inclusion or exclusion<br />

of certain groups have been going on for well over<br />

a century. In recent years, the main arguments<br />

have been between taxonomists striving <strong>to</strong>wards<br />

a phylogenetic definition based especially on the<br />

similarity of relevant DNA sequences, and others<br />

who take a biological approach <strong>to</strong> the subject and<br />

regard fungi as organisms sharing all or many key<br />

ecological or physiological characteristics the<br />

‘union of fungi’ (Barr, 1992). Being interested<br />

mainly in the way fungi function in nature and in<br />

the labora<strong>to</strong>ry, we take the latter approach and<br />

include several groups in this book which are now<br />

known <strong>to</strong> have arisen independently of the monophyletic<br />

‘true fungi’ (Eumycota) and have been<br />

placed outside them in recent classification<br />

schemes (see Fig. 1.25). The most important<br />

of these ‘pseudofungi’ are the Oomycota<br />

(see Chapter 5). Based on their lifestyle, fungi<br />

may be circumscribed by the following set of<br />

characteristics (modified from Ainsworth, 1973):<br />

1. Nutrition. Heterotrophic (lacking pho<strong>to</strong>synthesis),<br />

feeding by absorption rather than<br />

ingestion.<br />

2. Vegetative state. On or in the substratum,<br />

typically as a non-motile mycelium of<br />

hyphae showing internal pro<strong>to</strong>plasmic<br />

streaming. Motile reproductive states may<br />

occur.<br />

3. Cell wall. Typically present, usually based on<br />

glucans and chitin, rarely on glucans and<br />

cellulose (Oomycota).<br />

4. Nuclear status. Eukaryotic, uni- or multinucleate,<br />

the thallus being homo- or heterokaryotic,<br />

haploid, dikaryotic or diploid, the<br />

latter usually of short duration (but exceptions<br />

are known from several taxonomic<br />

groups).<br />

5. Life cycle. Simple or, more usually, complex.<br />

6. Reproduction. The following reproductive<br />

events may occur: sexual (i.e. nuclear<br />

fusion and meiosis) and/or parasexual<br />

(i.e. involving nuclear fusion followed by<br />

gradual de-diploidization) and/or asexual<br />

(i.e. purely mi<strong>to</strong>tic nuclear division).<br />

7. Propagules. These are typically microscopically<br />

small spores produced in high numbers.<br />

Motile spores are confined <strong>to</strong> certain<br />

groups.<br />

8. Sporocarps. Microscopic or macroscopic and<br />

showing characteristic shapes but only<br />

limited tissue differentiation.<br />

9. Habitat. Ubiqui<strong>to</strong>us in terrestrial and freshwater<br />

habitats, less so in the marine<br />

environment.<br />

10. Ecology. Important ecological roles as saprotrophs,<br />

mutualistic symbionts, parasites, or<br />

hyperparasites.<br />

11. Distribution. Cosmopolitan.


2 INTRODUCTION<br />

With pho<strong>to</strong>synthetic pigments being absent,<br />

fungi have a heterotrophic mode of nutrition.<br />

In contrast <strong>to</strong> animals which typically feed by<br />

ingestion, fungi obtain their nutrients by extracellular<br />

digestion due <strong>to</strong> the activity of secreted<br />

enzymes, followed by absorption of the solubilized<br />

breakdown products. The combination of<br />

extracellular digestion and absorption can be<br />

seen as the ultimate determinant of the fungal<br />

lifestyle. In the course of evolution, fungi have<br />

conquered an as<strong>to</strong>nishingly wide range of habitats,<br />

fulfilling important roles in diverse ecosystems<br />

(Dix & Webster, 1995). The conquest of new,<br />

often patchy resources is greatly facilitated by<br />

the production of numerous small spores rather<br />

than a few large propagules, whereas the<br />

colonization of a food source, once reached, is<br />

achieved most efficiently by growth as a system<br />

of branching tubes, the hyphae (Figs. 1.1a,b),<br />

which <strong>to</strong>gether make up the mycelium.<br />

Hyphae are generally quite uniform in different<br />

taxonomic groups of fungi. One of the few<br />

features of distinction that they do offer is the<br />

presence or absence of cross-walls or septa. The<br />

Oomycota and Zygomycota generally have aseptate<br />

hyphae in which the nuclei lie in a common<br />

mass of cy<strong>to</strong>plasm (Fig. 1.1a). Such a condition is<br />

described as coenocytic (Gr. koinos ¼ shared, in<br />

common; ky<strong>to</strong>s ¼ a hollow vessel, here meaning<br />

cell). In contrast, Asco- and Basidiomycota and<br />

their associated asexual states generally have<br />

septate hyphae (Fig. 1.1b) in which each segment<br />

contains one, two or more nuclei. If the nuclei<br />

are genetically identical, as in a mycelium<br />

derived from a single uninucleate spore, the<br />

mycelium is said <strong>to</strong> be homokaryotic, but where<br />

Fig1.1 Various growth forms of fungi. (a) Aseptate hypha of Mucor mucedo (Zygomycota).The hypha branches <strong>to</strong> form a mycelium.<br />

(b) Septate branched hypha of Trichoderma viride (Ascomycota). Septa are indicated by arrows. (c) Yeast cells of Schizosaccharomyces<br />

pombe (Ascomycota) dividing by binary fission. (d) Yeast cells of Dioszegia takashimae (Basidiomycota) dividing by budding.<br />

(e) Pseudohypha of Candida parapsilosis (Ascomycota), which is regarded as an intermediate stage between yeast cells and true<br />

hyphae. (f) Thallus of Rhizophlyctisrosea (Chytridiomycota) from which a system of branching rhizoids extends in<strong>to</strong> the substrate.<br />

(g) Plasmodia of Plasmodiophora brassicae (Plasmodiophoromycota) inside cabbage root cells. Scale bar ¼ 20 mm(a,b,f,g)or<br />

10 mm(c e).


PHYSIOLOGYOF THE GROWING HYPHA<br />

3<br />

a cell or mycelium contains nuclei of different<br />

genotype, e.g. as a result of fusion (anas<strong>to</strong>mosis)<br />

of genetically different hyphae, it is said <strong>to</strong> be<br />

heterokaryotic. A special condition is found in<br />

the mycelium of many Basidiomycota in which<br />

each cell contains two genetically distinct nuclei.<br />

This condition is dikaryotic, <strong>to</strong> distinguish it<br />

from mycelia which are monokaryotic. It should<br />

be noted that septa, where present, are usually<br />

perforated and allow for the exchange of<br />

cy<strong>to</strong>plasm or organelles.<br />

Not all fungi grow as hyphae. Some grow<br />

as discrete yeast cells which divide by fission<br />

(Fig. 1.1c) or, more frequently, budding (Fig. 1.1d).<br />

Yeasts are common, especially in situations<br />

where efficient penetration of the substratum<br />

is not required, e.g. on plant surfaces or in the<br />

digestive tracts of animals (Carlile, 1995). A few<br />

species, including certain pathogens of humans<br />

and animals, are dimorphic, i.e. capable of<br />

switching between hyphal and yeast-like growth<br />

forms (Gow, 1995). Intermediate stages between<br />

yeast cells and true hyphae also occur and are<br />

termed pseudohyphae (Fig. 1.1e). Some lower<br />

fungi grow as a thallus, i.e. a walled structure in<br />

which the pro<strong>to</strong>plasm is concentrated in one or<br />

more centres from which root-like branches<br />

(rhizoids) ramify (Fig. 1.1f). Certain obligately<br />

plant-pathogenic fungi and fungus-like organisms<br />

grow as a naked plasmodium (Fig. 1.1g),<br />

a uni- or multinucleate mass of pro<strong>to</strong>plasm<br />

not surrounded by a cell wall of its own, or as<br />

a pseudoplasmodium of amoeboid cells which<br />

retain their individual plasma membranes.<br />

However, by far the most important device<br />

which accounts for the typical biological features<br />

of fungi is the hypha (Bartnicki-Garcia, 1996),<br />

which therefore seems an appropriate starting<br />

point for an exploration of these organisms.<br />

1.2 Physiology of the<br />

growing hypha<br />

1.2.1 Polarity of the hypha<br />

By placing microscopic markers such as small<br />

glass beads beside a growing hypha, Reinhardt<br />

(1892) was able <strong>to</strong> show that cell wall extension,<br />

measured as an increase in the distance between<br />

two adjacent markers, occurred only at the<br />

extreme apex. Four years earlier, H. M. Ward<br />

(1888), in an equally simple experiment, had<br />

collected liquid droplets from the apex of hyphae<br />

of Botrytis cinerea and found that these ‘fermentdrops’<br />

were capable of degrading plant cell walls.<br />

Thus, the two fundamental properties of the vegetative<br />

fungal hypha the polarity of both growth<br />

and secretion of degradative enzymes have<br />

been known for over a century. Numerous studies<br />

have subsequently confirmed that ‘the key <strong>to</strong> the<br />

fungal hypha lies in the apex’ (Robertson, 1965),<br />

although the detailed mechanisms determining<br />

hyphal polarity are still obscure.<br />

Ultrastructural studies have shown that many<br />

organelles within the growing hyphal tip are<br />

distributed in steep gradients, as would be<br />

expected of a cell growing in a polarized<br />

mode (Girbardt, 1969; Howard, 1981). This<br />

is visible even with the light microscope by<br />

careful observation of an unstained hypha using<br />

phase-contrast optics (Reynaga-Peña et al., 1997),<br />

and more so with the aid of simple staining<br />

techniques (Figs. 1.2a d). The cy<strong>to</strong>plasm of<br />

the extreme apex is occupied almost exclusively<br />

by secre<strong>to</strong>ry vesicles and microvesicles<br />

(Figs. 1.2a, 1.3). In the higher fungi (Asco- and<br />

Basidiomycota), the former are arranged as a<br />

spherical shell around the latter, and the<br />

entire formation is called the Spitzenkörper or<br />

‘apical body’ (Fig. 1.4c; Bartnicki-Garcia, 1996).<br />

The Spitzenkörper may be seen in growing<br />

hyphae even with the light microscope. Hyphae<br />

of the Oomycota and some lower Eumycota<br />

(notably the Zygomycota) do not contain a<br />

recognizable Spitzenkörper, and the vesicles are<br />

instead distributed more loosely in the apical<br />

dome (Fig. 1.4a,b). Hyphal growth can be simulated<br />

by means of computer models based on the<br />

assumption that the emission of secre<strong>to</strong>ry vesicles<br />

is coordinated by a ‘vesicle supply centre’,<br />

regarded as the mathematical equivalent of the<br />

Spitzenkörper in higher fungi. By modifying<br />

certain parameters, it is even possible<br />

<strong>to</strong> generate the somewhat more pointed apex<br />

often found in hyphae of Oomycota and<br />

Zygomycota (Figs. 1.4a,b; Diéguez-Uribeondo<br />

et al., 2004).


4 INTRODUCTION<br />

Fig1.3 Transmission electron microscopy of a hyphal tip of<br />

Fusarium acuminatum preserved by the freeze-substitution<br />

method <strong>to</strong> reveal ultrastructural details.The vesicles of the<br />

Spitzenko«rper as well as mi<strong>to</strong>chondria (dark elongated<br />

organelles), a Golgi-like element (G) and microtubules<br />

(arrows) are visible. Microtubules are closely associated with<br />

mi<strong>to</strong>chondria. Reproduced from Howard and Aist (1980), by<br />

copyright permission of The Rockefeller University Press.<br />

Fig1.2 The organization of vegetative hyphae as seen by light microscopy. (a) Growing hypha of Galac<strong>to</strong>myces candidus showing the<br />

transition from dense apical <strong>to</strong> vacuolate basal cy<strong>to</strong>plasm.Tubular vacuolar continuities are also visible. (b e) His<strong>to</strong>chemistry in<br />

Botrytiscinerea. (b) Tetrazolium staining for mi<strong>to</strong>chondrial succinate dehydrogenase.The mi<strong>to</strong>chondria appear as dark filamen<strong>to</strong>us<br />

structures in subapical and maturing regions. (c) Staining of the same hypha for nuclei with the fluorescent DNA-binding dye DAPI.<br />

The apical cell contains numerous nuclei. (d) Staining of acid phosphatase activity using the Gomori lead-salt method with a fixed<br />

hypha. Enzyme activity is localized both in the secre<strong>to</strong>ry vesicles forming the Spitzenko«rper, and in vacuoles. (e) Uptake of Neutral<br />

Red in<strong>to</strong> vacuoles in a mature hyphal segment. All images <strong>to</strong> same scale.


PHYSIOLOGYOF THE GROWING HYPHA<br />

5<br />

Fig1.4 Schematic drawings of the arrangement of vesicles<br />

in growing hyphal tips. Secre<strong>to</strong>ry vesicles are visible in all<br />

hyphal tips, but the smaller microvesicles (chi<strong>to</strong>somes) are<br />

prominent only in Asco- and Basidiomycota and contribute<br />

<strong>to</strong> the Spitzenko«rper morphology of the vesicle cluster.<br />

(a) Oomycota. (b) Zygomycota. (c) Ascomycota and<br />

Basidiomycota.<br />

A little behind the apical dome, a region of<br />

intense biosynthetic activity and energy generation<br />

is indicated by parallel sheets of endoplasmic<br />

reticulum and an abundance of<br />

mi<strong>to</strong>chondria (Figs. 1.2b, 1.3). The first nuclei<br />

usually appear just behind the biosynthetic zone<br />

(Fig. 1.2c), followed ultimately by a system of<br />

ever-enlarging vacuoles (Fig. 1.2d). These may fill<br />

almost the entire volume of mature hyphal<br />

regions, making them appear empty when<br />

viewed with the light microscope.<br />

1.2.2 Architecture of the fungal cell wall<br />

Although the chemical composition of cell walls<br />

can vary considerably between and within<br />

different groups of fungi (Table 1.1), the basic<br />

design seems <strong>to</strong> be universal. It consists of<br />

a structural scaffold of fibres which are crosslinked,<br />

and a matrix of gel-like or crystalline<br />

material (Hunsley & Burnett, 1970; Ruiz-Herrera,<br />

1992; Sentandreu et al., 1994). The degree of<br />

cross-linking will determine the plasticity (extensibility)<br />

of the wall, whereas the pore size<br />

(permeability) is a property of the wall matrix.<br />

The scaffold forms the inner layer of the wall and<br />

the matrix is found predominantly in the outer<br />

layer (de Nobel et al., 2001).<br />

In the Ascomycota and Basidiomycota, the<br />

fibres are chitin microfibrils, i.e. bundles of<br />

linear b-(1,4)-linked N-acetylglucosamine chains<br />

(Fig. 1.5), which are synthesized at the plasma<br />

membrane and extruded in<strong>to</strong> the growing<br />

(‘nascent’) cell wall around the apical dome.<br />

The cell wall becomes rigid only after the<br />

microfibrils have been fixed in place by crosslinking.<br />

These cross-links consist of highly<br />

branched glucans (glucose polymers), especially<br />

those in which the glucose moieties are linked by<br />

b-(1,3)- and b-(1,6)-bonds (Suarit et al., 1988;<br />

Wessels et al., 1990; Sietsma & Wessels, 1994).<br />

Such b-glucans are typically insoluble in alkaline<br />

solutions (1 M KOH). In contrast, the alkalisoluble<br />

glucan fraction contains mainly a-(1,3)-<br />

and/or a-(1,4)-linked branched or unbranched<br />

chains (Wessels et al., 1972; Bobbitt & Nordin,<br />

1982) and does not perform a structural role<br />

but instead contributes significantly <strong>to</strong> the<br />

cell wall matrix (Sietsma & Wessels, 1994).<br />

Proteins represent the third important chemical<br />

Table 1.1. The chemical composition of cell walls of selected groups of fungi (dry weight of <strong>to</strong>tal cell wall<br />

fraction, in per cent). Data adapted from Ruiz-Herrera (1992) and Griffin (1994).<br />

Group Example Chitin Cellulose Glucans Protein Lipid<br />

Oomycota Phy<strong>to</strong>phthora 0 25 65 4 2<br />

Chytridiomycota Allomyces 58 0 16 10 ?<br />

Zygomycota Mucor 9 0 44 6 8<br />

Ascomycota Saccharomyces 1 0 60 13 8<br />

Fusarium 39 0 29 7 6<br />

Basidiomycota Schizophyllum 5 0 81 2 ?<br />

Coprinus 33 0 50 10 ?<br />

Mainly chi<strong>to</strong>san.


6 INTRODUCTION<br />

Fig1.5 Structural formulae of the principal fibrous<br />

components of fungal cell walls.<br />

constituent of fungal cell walls. In addition <strong>to</strong><br />

enzymes involved in cell wall synthesis or lysis,<br />

or in extracellular digestion, there are also<br />

structural proteins. Many cell wall proteins are<br />

modified by glycosylation, i.e. the attachment of<br />

oligosaccharide chains <strong>to</strong> the polypeptide. The<br />

degree of glycosylation can be very high, especially<br />

in the yeast Saccharomyces cerevisiae, where<br />

up <strong>to</strong> 90% of the molecular weight of an<br />

extracellular protein may be contributed by its<br />

glycosylation chains (van Rinsum et al., 1991).<br />

Since mannose is the main component, such<br />

proteins are often called mannoproteins or<br />

mannans. In S. cerevisiae, the pore size of the<br />

cell wall is determined not by matrix glucans but<br />

by mannoproteins located close <strong>to</strong> the external<br />

wall surface (Zlotnik et al., 1984). Proteins<br />

exposed at the cell wall surface can also<br />

determine surface properties such as adhesion<br />

and recognition (Cormack et al., 1999). Structural<br />

proteins often contain a glycosylphosphatidylinosi<strong>to</strong>l<br />

anchor by which they are attached <strong>to</strong> the<br />

lumen of the rough endoplasmic reticulum (ER)<br />

and later <strong>to</strong> the external plasma membrane<br />

surface, or a modified anchor which covalently<br />

binds them <strong>to</strong> the b-(1,6)-glucan fraction of the<br />

cell wall (Kollár et al., 1997; de Nobel et al., 2001).<br />

In the Zygomycota, the chitin fibres are<br />

modified after their synthesis by partial or<br />

complete deacetylation <strong>to</strong> produce poly-b-(1,4)-<br />

glucosamine, which is called chi<strong>to</strong>san (Fig. 1.5)<br />

(Calvo-Mendez & Ruiz-Herrera, 1987). Chi<strong>to</strong>san<br />

fibres are cross-linked by polysaccharides containing<br />

glucuronic acid and various neutral<br />

sugars (Datema et al., 1977). The cell wall<br />

matrix comprises glucans and proteins, as it<br />

does in members of the other fungal groups.<br />

One traditional feature <strong>to</strong> distinguish the<br />

Oomycota from the ‘true fungi’ (Eumycota) has<br />

been the absence of chitin from their cell walls<br />

(Wessels & Sietsma, 1981), even though chitin is<br />

now known <strong>to</strong> be produced by certain species of<br />

Oomycota under certain conditions (Gay et al.,<br />

1993). By and large, however, in Oomycota, the<br />

structural role of chitin is filled by cellulose, an<br />

aggregate of linear b-(1,4)-glucan chains (Fig. 1.5).<br />

As in many other fungi, the fibres thus produced<br />

are cross-linked by an alkali-insoluble glucan<br />

containing b-(1,3)- and b-(1,6)-linkages. In addition<br />

<strong>to</strong> proteins, the main matrix component<br />

appears <strong>to</strong> be an alkali-soluble b-(1,3)-glucan<br />

(Wessels & Sietsma, 1981).<br />

1.2.3 Synthesis of the cell wall<br />

The synthesis of chitin is mediated by specialized<br />

organelles termed chi<strong>to</strong>somes (Bartnicki-<br />

Garcia et al., 1979; Sentandreu et al., 1994) in<br />

which inactive chitin synthases are delivered <strong>to</strong><br />

the apical plasma membrane and become activated<br />

upon contact with the lipid bilayer<br />

(Montgomery & Gooday, 1985). Microvesicles,<br />

visible especially in the core region of the<br />

Spitzenkörper, are likely <strong>to</strong> be the ultrastructural<br />

manifestation of chi<strong>to</strong>somes (Fig. 1.6). In<br />

contrast, structural proteins and enzymes travel<br />

<strong>to</strong>gether in the larger secre<strong>to</strong>ry vesicles and<br />

are discharged in<strong>to</strong> the environment when<br />

the vesicles fuse with the plasma membrane


PHYSIOLOGYOF THE GROWING HYPHA<br />

7<br />

Fig1.6 The Spitzenko«rper of Botrytis cinerea which is<br />

differentiated in<strong>to</strong> an electron-dense core consisting of<br />

microvesicles (chi<strong>to</strong>somes) and an outer region made up of<br />

larger secre<strong>to</strong>ry vesicles, some of which are located close<br />

<strong>to</strong> the plasma membrane. Reprinted from Weber and Pitt<br />

(2001), with permission from Elsevier.<br />

(Fig. 1.6). Whereas most proteins are fully<br />

functional by the time they traverse the plasma<br />

membrane (see p. 10), the glucans are secreted by<br />

secre<strong>to</strong>ry vesicles as partly formed precursors<br />

(Wessels, 1993a) and undergo further polymerization<br />

in the nascent cell wall, or they are<br />

synthesized entirely at the plasma membrane<br />

(Sentandreu et al., 1994; de Nobel et al., 2001).<br />

Cross-linking of glucans with other components<br />

of the cell wall takes place after extrusion in<strong>to</strong><br />

the cell wall (Kollár et al., 1997; de Nobel et al.,<br />

2001).<br />

Wessels et al. (1990) have provided experimental<br />

evidence <strong>to</strong> support a model for<br />

cell wall synthesis in Schizophyllum commune<br />

(Basidiomycota). The individual linear b-(1,4)-Nacetylglucosamine<br />

chains extruded from the<br />

plasma membrane are capable of undergoing<br />

self-assembly in<strong>to</strong> chitin microfibrils, but this is<br />

subject <strong>to</strong> a certain delay during which crosslinking<br />

with glucans must occur. The glucans,<br />

in turn, become alkali-insoluble only after they<br />

have become covalently linked <strong>to</strong> chitin. Once<br />

the structural scaffold is in place, the wall matrix<br />

can be assembled. Wessels (1997) suggested that<br />

hyphal growth occurs as the result of a continuously<br />

replenished supply of soft wall material at<br />

the apex, but there is good evidence that the<br />

softness of the apical cell wall is also influenced<br />

by the activity of wall-lytic enzymes such as<br />

chitinases or glucanases (Fontaine et al., 1997;<br />

Horsch et al., 1997). Further, when certain<br />

Oomycota grow under conditions of hyperosmotic<br />

stress, their cell wall is measurably softer<br />

due <strong>to</strong> the secretion of an endo-b-(1,4)-glucanase,<br />

thus permitting continued growth when the<br />

turgor pressure is reduced or even absent<br />

(Money, 1994; Money & Hill, 1997). Since, in<br />

higher Eumycota, both cell wall material and<br />

synthetic as well as lytic enzymes are secreted<br />

<strong>to</strong>gether by the vesicles of the Spitzenkörper,<br />

the appearance, position and movement of<br />

this structure should influence the direction<br />

and speed of apical growth directly. This has<br />

indeed been shown <strong>to</strong> be the case (López-Franco<br />

et al., 1995; Bartnicki-Garcia, 1996; Riquelme<br />

et al., 1998).<br />

Of course, cell wall-lytic enzymes are also<br />

necessary for the formation of hyphal branches,<br />

which usually arise by a localized weakening of<br />

the mature, fully polymerized cell wall. An<br />

endo-b-(1,4)-glucanase has also been shown <strong>to</strong> be<br />

involved in softening the mature regions of<br />

hyphae in the growing stipes of Coprinus fruit<br />

bodies, thus permitting intercalary hyphal<br />

extension (Kamada, 1994). Indeed, the expansion


8 INTRODUCTION<br />

of mushroom-type fruit bodies in general seems<br />

<strong>to</strong> be based mainly on non-apical extension<br />

of existing hyphae (see p. 22), which is a<br />

rare exception <strong>to</strong> the rule of apical growth in<br />

fungi.<br />

The properties of the cell wall depend in<br />

many ways on the environment in which the<br />

hypha grows. Thus, when Schizophyllum commune<br />

is grown in liquid submerged culture, a significant<br />

part of the b-glucan fraction may diffuse<br />

in<strong>to</strong> the liquid medium before it is captured by<br />

the cell wall, giving rise <strong>to</strong> mucilage (Sietsma<br />

et al., 1977). In addition <strong>to</strong> causing problems<br />

when growing fungi in liquid culture for experimental<br />

purposes, mucilage may cause economic<br />

losses when released by Botrytis cinerea in grapes<br />

<strong>to</strong> be used for wine production (Dubourdieu<br />

et al., 1978a). On the other hand, secreted<br />

polysaccharides, especially of Basidiomycota,<br />

may have interesting medicinal properties and<br />

are being promoted as anti-tumour medication<br />

both in conventional and in alternative medicine<br />

(Wasser, 2002).<br />

Another difference between submerged and<br />

aerial hyphae is caused by the hydrophobins,<br />

which are structural cell wall proteins with<br />

specialized functions in physiology, morphogenesis<br />

and pathology (Wessels, 2000). Some hydrophobins<br />

are constitutively secreted by the hyphal<br />

apex. In submerged culture, they diffuse in<strong>to</strong> the<br />

medium as monomers, whereas they polymerize<br />

by hydrophobic interactions on the surface<br />

of hyphae exposed <strong>to</strong> air, thereby effectively<br />

impregnating them and rendering them<br />

hydrophobic (Wessels, 1997, 2000). When freezefractured<br />

hydrophobic surfaces of hyphae or<br />

spores are viewed with the transmission electron<br />

microscope, polymerized hydrophobins may be<br />

visible as patches of rodlets running in parallel<br />

<strong>to</strong> each other. Other hydrophobins are produced<br />

only at particular developmental stages and are<br />

involved in inducing morphogenetic changes of<br />

the hypha, leading, for example, <strong>to</strong> the formation<br />

of spores or infection structures, or aggregation<br />

of hyphae in<strong>to</strong> fruit bodies (Stringer et al.,<br />

1991; Wessels, 1997).<br />

Some fungi are wall-less during the assimilative<br />

stage of their life cycle. This is true especially<br />

of certain plant pathogens such as the<br />

Plasmodiophoromycota (Chapter 3), insect pathogens<br />

(En<strong>to</strong>mophthorales; p. 202) and some<br />

members of the Chytridiomycota (Chapter 6).<br />

Since their pro<strong>to</strong>plasts are in direct contact<br />

with the host cy<strong>to</strong>plasm, they are buffered<br />

against osmotic fluctuations. The motile spores<br />

(zoospores) of certain groups of fungi swim freely<br />

in water, and bursting due <strong>to</strong> osmotic inward<br />

movement of water is prevented by the constant<br />

activity of water-expulsion vacuoles.<br />

1.2.4 The cy<strong>to</strong>skele<strong>to</strong>n<br />

In contrast <strong>to</strong> the hyphae of certain Oomycota,<br />

which seem <strong>to</strong> grow even in the absence of<br />

measurable turgor pressure (Money & Hill, 1997),<br />

the hyphae of most fungi extend only when a<br />

threshold turgor pressure is exceeded. This can<br />

be generated even at a reduced external water<br />

potential by the accumulation of compatible<br />

solutes such as glycerol, manni<strong>to</strong>l or trehalose<br />

inside the hypha (Jennings, 1995). The correlation<br />

between turgor pressure and hyphal growth<br />

might be interpreted such that the former drives<br />

the latter, but this crude mechanism would lead<br />

<strong>to</strong> uncontrolled tip extension or even tip bursting.<br />

Further, when hyphal tips are made <strong>to</strong> burst<br />

by experimental manipulation, they often do<br />

so not at the extreme apex, but a little further<br />

behind (Sietsma & Wessels, 1994). It seems,<br />

therefore, that the soft wall at the apex is protected<br />

internally, and there is now good evidence<br />

that this is mediated by the cy<strong>to</strong>skele<strong>to</strong>n.<br />

Both main elements of the cy<strong>to</strong>skele<strong>to</strong>n,<br />

i.e. micro<strong>to</strong>bules (Figs. 1.7a,b) and actin filaments<br />

(Fig. 1.7c), are abundant in filamen<strong>to</strong>us fungi and<br />

yeasts (Heath, 1994, 1995a). Intermediate filaments,<br />

which fulfil skeletal roles in animal cells,<br />

are probably of lesser significance in fungi.<br />

Microtubules are typically orientated longitudinally<br />

relative <strong>to</strong> the hypha (Fig. 1.7a) and are<br />

involved in long-distance transport of organelles<br />

such as secre<strong>to</strong>ry vesicles (Fig. 1.7b; Seiler et al.,<br />

1997) or nuclei (Steinberg, 1998), and in the<br />

positioning of mi<strong>to</strong>chondria, nuclei or vacuoles<br />

(Howard & Aist, 1977; Steinberg et al., 1998). They<br />

therefore maintain the polarized distribution of<br />

many organelles in the hyphal tip.


PHYSIOLOGYOF THE GROWING HYPHA<br />

9<br />

Fig1.7 The cy<strong>to</strong>skele<strong>to</strong>n in fungi. (a) Microtubules in Rhizoc<strong>to</strong>nia solani (Basidiomycota) stained with an a-tubulin antibody.<br />

(b) Secre<strong>to</strong>ry vesicles (arrowheads) associated with a microtubule in Botrytis cinerea (Ascomycota). (c) The actin system of<br />

Saprolegnia ferax (Oomycota) stained with phalloidin rhodamine. Note the dense actin cap in growing hyphal tips. (a) reproduced<br />

from Bourett et al. (1998), with permission from Elsevier; original print kindly provided by R. J.Howard. (b) reproduced fromWeber<br />

and Pitt (2001), with permission from Elsevier. (c) reproduced from I.B. Heath (1987), by copyright permission of Wissenschaftliche<br />

Verlagsgesellschaft mbH, Stuttgart; original print kindly provided by I.B. Heath.<br />

Actin filaments are found in the centre of the<br />

Spitzenkörper, as discrete subapical patches, and<br />

as a cap lining the inside of the extreme hyphal<br />

apex (Heath, 1995a; Czymmek et al., 1996;<br />

Srinivasan et al., 1996). The apical actin cap is<br />

particularly pronounced in Oomycota such as<br />

Saprolegnia (Fig. 1.7c), and it now seems that the<br />

soft wall at the hyphal apex is actually being<br />

assembled on an internal scaffold consisting of<br />

actin and other structural proteins, such as<br />

spectrin (Heath, 1995b; Degousée et al., 2000).<br />

The rate of hyphal extension might be<br />

controlled, and bursting prevented, by the<br />

actin/spectrin cap being anchored <strong>to</strong> the rigid,<br />

subapical wall via rivet-like integrin attachments<br />

which traverse the membrane and might bind <strong>to</strong><br />

wall matrix proteins (Fig. 1.8; Kaminskyj &<br />

Heath, 1996; Heath, 2001). Indeed, in Saprolegnia<br />

the cy<strong>to</strong>skele<strong>to</strong>n is probably responsible for<br />

pushing the hyphal tip forward, at least in the<br />

absence of turgor (Money, 1997), although it<br />

probably has a restraining function under<br />

normal physiological conditions. Heath (1995b)<br />

has proposed an ingenious if speculative model<br />

<strong>to</strong> explain how the actin cap might regulate the<br />

rate of hyphal tip extension in the Oomycota.<br />

Stretch-activated channels selective for Ca 2þ ions<br />

are known <strong>to</strong> be concentrated in the apical<br />

plasma membrane of Saprolegnia (Garrill et al.,<br />

1993), and the fact that Ca 2þ ions cause contractions<br />

of actin filaments is also well known.<br />

A stretched plasma membrane will admit Ca 2þ<br />

ions in<strong>to</strong> the apical cy<strong>to</strong>plasm where they cause<br />

localized contractions of the actin cap, thereby<br />

reducing the rate of apical growth which leads <strong>to</strong><br />

closure of the stretch-activated Ca 2þ channels.<br />

Sequestration of Ca 2þ by various subapical<br />

organelles such as the ER or vacuoles lowers<br />

the concentration of free cy<strong>to</strong>plasmic Ca 2þ ,<br />

leading <strong>to</strong> a relaxation of the actin cap and of<br />

its restrictive effect on hyphal growth.<br />

In the Eumycota, there is only indirect<br />

evidence for a similar role of actin, integrin<br />

and other structural proteins in protecting the<br />

apex and restraining its extension (Degousée<br />

et al., 2000; Heath, 2001), and the details of


10 INTRODUCTION<br />

Fig1.8 Diagrammatic representation of the internal scaffold<br />

model of tip growth in fungi proposed by Heath (1995b).<br />

Secre<strong>to</strong>ry vesicles and chi<strong>to</strong>somes are transported along<br />

microtubules from their subapical sites of synthesis <strong>to</strong> the<br />

growing apex.The Spitzenko«rper forms around a cluster of<br />

actin filaments. An actin scaffold inside the extreme apex is<br />

linked <strong>to</strong> rivet-like integrin molecules which are anchored in<br />

the rigid subapical cell wall.The apex is further stabilized by<br />

spectrin molecules lining the cy<strong>to</strong>plasmic surface of the<br />

plasma membrane. Redrawn and modified from Weber and<br />

Pitt (2001).<br />

regulation are likely <strong>to</strong> be different. Whereas a<br />

tip-high Ca 2þ gradient is present and is required<br />

for growth, stretch-activated Ca 2þ channels<br />

are not, and the apical Ca 2þ seems <strong>to</strong> be<br />

of endogenous origin. Silverman-Gavrila and<br />

Lew (2001, 2002) have proposed that the signal<br />

molecule inosi<strong>to</strong>l-(1,4,5)-trisphosphate (IP 3 ),<br />

released by the action of a stretch-activated<br />

phospholipase C in the apical plasma membrane,<br />

acts on Ca 2þ -rich secre<strong>to</strong>ry vesicles in the<br />

Spitzenkörper region. These would release Ca 2þ<br />

from their lumen, leading <strong>to</strong> a contraction of the<br />

apical scaffold. As in the Oomycota, sequestration<br />

of Ca 2þ occurs subapically by the ER from<br />

which secre<strong>to</strong>ry vesicles are formed. These therefore<br />

act as Ca 2þ shuttles in the Eumycota<br />

(Torralba et al., 2001). Although hyphal tip<br />

growth appears <strong>to</strong> be a straightforward affair,<br />

none of the conflicting models accounts for<br />

all aspects of it. A good essay in hyphal tip<br />

diplomacy has been written by Bartnicki-Garcia<br />

(2002).<br />

Numerous inhibi<strong>to</strong>r studies have hinted at<br />

a role of the cy<strong>to</strong>skele<strong>to</strong>n in the transport of<br />

vesicles <strong>to</strong> the apex. Depolymerization of<br />

microtubules results in a disappearance of the<br />

Spitzenkörper, termination or at least severe<br />

reduction of apical growth and enzyme secretion,<br />

and an even redistribution of secre<strong>to</strong>ry<br />

vesicles and other organelles throughout the<br />

hypha (Howard & Aist, 1977; Rupeš et al., 1995;<br />

Horio & Oakley, 2005). In contrast, actin depolymerization<br />

leads <strong>to</strong> uncontrolled tip extension<br />

<strong>to</strong> form giant spheres (Srinivasan et al., 1996).<br />

Long-distance transport of secre<strong>to</strong>ry vesicles<br />

therefore seems <strong>to</strong> be brought about by microtubules,<br />

whereas the fine-tuning of vesicle fusion<br />

with the plasma membrane is controlled by actin<br />

(Fig. 1.8; Torralba et al., 1998). The integrity of the<br />

Spitzenkörper is maintained by an interplay<br />

between actin and tubulin. Not surprisingly,<br />

the yeast S. cerevisiae, which has a very short<br />

vesicle transport distance between the mother<br />

cell and the extending bud, reacts more sensitively<br />

<strong>to</strong> disruptions of the actin component<br />

than the microtubule component of its cy<strong>to</strong>skele<strong>to</strong>n;<br />

continued growth in the absence of<br />

the latter can be explained by Brownian motion<br />

of secre<strong>to</strong>ry vesicles (Govindan et al., 1995;<br />

Steinberg, 1998).<br />

1.2.5 Secretion and membrane traffic<br />

One of the most important ecological roles of<br />

fungi, that of decomposing dead plant matter,<br />

requires the secretion of large quantities of<br />

hydrolytic and oxidative enzymes in<strong>to</strong> the<br />

environment. In liquid culture under optimized<br />

experimental conditions, certain fungi


PHYSIOLOGYOF THE GROWING HYPHA<br />

11<br />

are capable of secreting more than 20 g of a<br />

single enzyme or enzyme group per litre culture<br />

broth within a few days’ growth (Sprey, 1988;<br />

Peberdy, 1994). Clearly, this aspect of fungal<br />

physiology holds considerable potential for<br />

biotechnological or pharmaceutical applications.<br />

However, for reasons not yet entirely unders<strong>to</strong>od,<br />

fungi often fail <strong>to</strong> secrete the heterologous<br />

proteins of introduced genes of commercial<br />

interest <strong>to</strong> the same high level as their own<br />

proteins (Gwynne, 1992). There are still<br />

great deficits in our understanding of the<br />

fundamental mechanisms of the secre<strong>to</strong>ry<br />

route in filamen<strong>to</strong>us fungi, although much is<br />

known in the yeast S. cerevisiae. An overview is<br />

given in Fig. 1.10.<br />

As in other eukaryotes, the secre<strong>to</strong>ry route in<br />

fungi begins in the ER. Ribosomes loaded with<br />

a suitable messenger RNA dock on<strong>to</strong> the<br />

ER membrane and translate the polypeptide<br />

product which enters the ER lumen during<br />

its synthesis unless specific internal signal<br />

sequences cause it <strong>to</strong> be retained in the ER<br />

membrane. As soon as the protein is in contact<br />

with the ER lumen, oligosaccharide chains may<br />

be added on<strong>to</strong> selected amino acids. These<br />

glycosylation chains are subject <strong>to</strong> successive<br />

modification steps as the protein traverses the<br />

secre<strong>to</strong>ry route, whereby the chains in S. cerevisiae<br />

become considerably larger than those in most<br />

filamen<strong>to</strong>us fungi (Maras et al., 1997; Gemmill &<br />

Trimble, 1999). Paradoxically, even though filamen<strong>to</strong>us<br />

fungi possess such powerful secre<strong>to</strong>ry<br />

systems, morphologically recognizable Golgi<br />

stacks have not generally been observed except<br />

for the Oomycota, Plasmodiophoromycota and<br />

related groups (Grove et al., 1968; Beakes &<br />

Glockling, 1998). In all other fungi, the Golgi<br />

apparatus seems <strong>to</strong> be much reduced <strong>to</strong> single<br />

cisternae (Howard, 1981; see Fig. 1.3), with<br />

images of fully fledged Golgi stacks only<br />

published occasionally (see e.g. Fig. 10.1). In<br />

S. cerevisiae and probably also in filamen<strong>to</strong>us<br />

fungi, the transport of proteins from the ER<br />

<strong>to</strong> the Golgi system occurs via vesicular<br />

carriers (Schekman, 1992), although continuous<br />

membrane flow is also possible (see p. 272).<br />

Membrane lipids seem <strong>to</strong> be recycled <strong>to</strong> the ER<br />

by a different mechanism relying on tubular<br />

continuities (Rupeš et al., 1995; Akashi et al.,<br />

1997).<br />

In the Golgi system, proteins are subjected <strong>to</strong><br />

stepwise further modifications (Graham & Emr,<br />

1991), and proteins destined for the vacuolar<br />

system are separated from those bound for<br />

secretion (Seeger & Payne, 1992). Both destinations<br />

are probably reached by vesicular carriers,<br />

the secre<strong>to</strong>ry vesicles moving along microtubules<br />

<strong>to</strong> reach the growing hyphal apex (Fig. 1.7b),<br />

which is the site for secretion of extracellular<br />

enzymes as well as new cell wall material<br />

(Peberdy, 1994). Collinge and Trinci (1974)<br />

estimated that 38 000 secre<strong>to</strong>ry vesicles per<br />

minute fuse with the plasma membrane of<br />

a single growing hypha of Neurospora crassa.<br />

Microvesicles (chi<strong>to</strong>somes) probably arise from a<br />

discrete population of Golgi cisternae (Howard,<br />

1981).<br />

There is mounting evidence that fungi, like<br />

most eukaryotes, are capable of performing<br />

endocy<strong>to</strong>sis by the inward budding of the<br />

plasma membrane at subapical locations.<br />

Endocy<strong>to</strong>sis may be necessary <strong>to</strong> retrieve<br />

membrane material in excess of that which is<br />

required for extension at the growing apex, i.e.<br />

endocy<strong>to</strong>sis and exocy<strong>to</strong>sis may be coupled<br />

(Steinberg & Fuchs, 2004). The prime destination<br />

of endocy<strong>to</strong>sed membrane material or vital<br />

stains is the vacuole (Vida & Emr, 1995; Fischer-<br />

Par<strong>to</strong>n et al., 2000; Weber, 2002). In fungi, large<br />

vacuoles (Figs. 1.2e, 1.9) represent the main<br />

element of the lytic system and are the sink<br />

not only for endocy<strong>to</strong>sed material but also<br />

for au<strong>to</strong>phagocy<strong>to</strong>sis, i.e. the sequestration<br />

and degradation of organelles or cy<strong>to</strong>plasm.<br />

Au<strong>to</strong>phagocy<strong>to</strong>sis is especially prominent under<br />

starvation conditions (Takeshige et al., 1992).<br />

Careful ultrastructural studies have revealed<br />

that adjacent vacuoles may be linked by thin<br />

membranous tubes, thereby providing a potential<br />

means of transport (Rees et al., 1994). These<br />

tubes can extend even through the septal<br />

pores and show peristaltic movement, possibly<br />

explaining why especially mycorrhizal fungi are<br />

capable of rapid translocation of solutes over<br />

long hyphal distances (Fig. 1.9; Cole et al., 1998;<br />

Ashford et al., 2001).


12 INTRODUCTION<br />

Fig1.9 Tubular continuities linking adjacent vacuoles of Pisolithus tinc<strong>to</strong>rius. (a) Light micrograph of the vacuolar system of Pisolithus<br />

tinc<strong>to</strong>rius stained with a fluorescent dye. (b) TEM image of a freeze-substituted hypha. Reproduced from Ashford et al.(2001),with<br />

kind permission of Springer Science and Business Media.Original images kindly provided by A.E. Ashford.<br />

1.2.6 Nutrient uptake<br />

One of the hallmarks of fungi is their ability <strong>to</strong><br />

take up organic or inorganic solutes from<br />

extremely dilute solutions in the environment,<br />

accumulating them 1000-fold or more against<br />

their concentration gradient (Griffin, 1994). The<br />

main barrier <strong>to</strong> the movement of water-soluble<br />

substances in<strong>to</strong> the cell is the lipid bilayer of<br />

the plasma membrane. Uptake is mediated by<br />

proteinaceous pores in the plasma membrane<br />

which are always selective for particular solutes.<br />

The pores are termed channels (system I) if they<br />

facilitate the diffusion of a solute following its<br />

concentration gradient whilst they are called<br />

porters (system II) if they use metabolic energy<br />

<strong>to</strong> accumulate the solute across the plasma<br />

membrane against its gradient (Harold, 1994).<br />

<strong>Fungi</strong> often possess one channel and one porter<br />

for a given solute. The high-affinity porter system<br />

is repressed at high external solute concentrations<br />

such as those found in most labora<strong>to</strong>ry<br />

media (Scarborough, 1970; Sanders, 1988).<br />

In nature, however, the concentration of<br />

nutrients is often so low that the porter systems<br />

are active. Porters do not directly convert<br />

metabolic energy (ATP) in<strong>to</strong> the uptake of<br />

solutes; rather, ATP is hydrolysed by ATPases<br />

which pump pro<strong>to</strong>ns (H þ ) <strong>to</strong> the outside of<br />

the plasma membrane, thus establishing a<br />

transmembrane pH gradient (acid outside). It<br />

has been estimated that one-third of the <strong>to</strong>tal<br />

cellular ATP is used for the establishment of<br />

the transmembrane H þ gradient (Gradmann<br />

et al., 1978). The inward movement of H þ following<br />

its electrochemical gradient is harnessed<br />

by the porters for solute uptake by means of<br />

solute porter H þ complexes (Slayman &<br />

Slayman, 1974; Slayman, 1987; Garrill, 1995).<br />

Different types of porter exist, depending on<br />

the charge of the desired solute. Uniport and<br />

symport carriers couple the inward movement<br />

of H þ with the uptake of uncharged or negatively<br />

charged solutes, respectively, whereas antiports<br />

harness the outward diffusion of cations such as<br />

K þ for the uptake of other positively charged<br />

solutes. Charge imbalances can be rectified by<br />

the selective opening of K þ channels. Porters<br />

have been described for NH þ 4 ,NO 3 , amino acids,<br />

hexoses, orthophosphate and other solutes<br />

(Garrill, 1995; Jennings, 1995).<br />

The ATPases fuelling active uptake mechanisms<br />

are located in subapical or mature regions<br />

of the plasma membrane, whereas the porter<br />

systems are typically situated in the apical<br />

membrane (Harold, 1994), closest <strong>to</strong> the site<br />

where the solutes may be released by the activity<br />

of extracellular enzymes. Thus, mature hyphal<br />

segments make a substantial direct contribution


PHYSIOLOGYOF THE GROWING HYPHA<br />

13<br />

(Fig. 1.11), which was at one time thought <strong>to</strong> be<br />

a causal fac<strong>to</strong>r of hyphal tip polarity but is now<br />

regarded as a consequence of it (Harold, 1994).<br />

Pro<strong>to</strong>n pumps fuelled by ATP are prominent<br />

also in the vacuolar membrane, the <strong>to</strong>noplast<br />

(Fig. 1.11), and their activity acidifies the vacuolar<br />

lumen (Klionsky et al., 1990). The principle of<br />

pro<strong>to</strong>n-coupled solute transport is utilized by<br />

the vacuole <strong>to</strong> fulfil its role as a system for the<br />

s<strong>to</strong>rage of nutrients, for example phosphate<br />

(Cramer & Davis, 1984) or amino acids such as<br />

arginine (Keenan & Weiss, 1997), or for the<br />

removal of <strong>to</strong>xic compounds from the cy<strong>to</strong>plasm,<br />

e.g. Ca 2þ or heavy metal ions (Cornelius &<br />

Nakashima, 1987).<br />

Fig1.10 Schematic summary of the pathways of membrane<br />

flow in a growing hypha. Secre<strong>to</strong>ry proteins (), vacuolar<br />

luminal proteins (), membrane-bound proteins ( ),<br />

endocy<strong>to</strong>sed (g) and au<strong>to</strong>phagocy<strong>to</strong>sed (m) material is<br />

indicated, as are vacuolar degradation products (). Redrawn<br />

and modified from Weber (2002).<br />

<strong>to</strong> the growth of the hypha at its tip. The spatial<br />

separation of H þ expulsion and re-entry generates<br />

an external electric field carried by pro<strong>to</strong>ns<br />

1.2.7 Hyphal branching<br />

Assimilative hyphae of most fungi grow monopodially<br />

by a main axis (leading hypha) capable<br />

of potentially unlimited apical growth. Branches<br />

arise at some distance behind the apex, suggesting<br />

some form of apical dominance, i.e. the<br />

presence of a growing apex inhibits the development<br />

of lateral branches close <strong>to</strong> it. Dicho<strong>to</strong>mous<br />

branching is rare, but does occur in Allomyces<br />

(see Fig. 6.20d) and Galac<strong>to</strong>myces geotrichum.<br />

In septate fungi, branches are often located<br />

immediately behind a septum. Branches usually<br />

arise singly in vegetative hyphae, although<br />

whorls of branches (i.e. branches arising near a<br />

common point) occur in reproductive structures.<br />

Branching may thus be under genetic or external<br />

control (Burnett, 1976). An even spacing between<br />

vegetative hyphae results from a combination of<br />

chemotropic growth <strong>to</strong>wards a source of diffusible<br />

nutrients, and growth away from staling<br />

products secreted by other hyphae which have<br />

colonized a substratum. The circular appearance<br />

of fungal colonies in Petri dish cultures<br />

arises because certain lateral branches grow out<br />

and fill the space between the leading radial<br />

branches, keeping pace with their rate of growth.<br />

This invasive growth is the most efficient way <strong>to</strong><br />

spread throughout a substratum. In nature,<br />

it may be obvious even <strong>to</strong> the naked eye,<br />

for example, in the shape of fairy rings (see<br />

Figs. 19.18a,b).


14 INTRODUCTION<br />

Fig1.11 Ion fluxes in a growing hypha.The<br />

pro<strong>to</strong>n (H þ ) gradient across the plasma<br />

membrane is generated by subapical<br />

ATP-driven expulsion of pro<strong>to</strong>ns. It is used<br />

for the active uptake of nutrients by porters.<br />

Channels also exist for most of the nutrients<br />

but are not shown here, except for the K þ<br />

channel which operates <strong>to</strong> compensate for<br />

charge imbalances. Dotted arrows<br />

indicate movement of a solute against its<br />

concentration gradient; solid arrows<br />

indicate movement from concentrated <strong>to</strong><br />

dilute.For details, see Garrill (1995).<br />

1.3 Hyphal aggregates<br />

Whereas plants and animals form genuine<br />

tissues by their ability <strong>to</strong> perform cell divisions<br />

in all directions, fungi are limited by their<br />

growth as one-dimensional hyphae. None the<br />

less, fungi are capable of producing complex<br />

and characteristic multicellular structures which<br />

resemble the tissues of other eukaryotes. This<br />

must be controlled by the positioning, growth<br />

rate and growth direction of individual hyphal<br />

branches (Moore, 1994). Further, instead of<br />

spacing themselves apart as during invasive<br />

growth, hyphae must be made <strong>to</strong> aggregate.<br />

Very little is known about the signalling events


HYPHAL AGGREGATES<br />

15<br />

leading <strong>to</strong> the synchronized growth of groups<br />

of hyphae. However, it may be speculated that<br />

the diffusion of signalling molecules takes place<br />

between adjacent hyphae, i.e. that a given hypha<br />

is able <strong>to</strong> influence the gene expression of adjacent<br />

hyphae by secreting chemical messengers.<br />

This may be facilitated by an extrahyphal glucan<br />

matrix within which aggregating hyphae are<br />

typically embedded (Moore, 1994). Such matrices<br />

have been found in rhizomorphs (Rayner et al.,<br />

1985), sclerotia (Fig. 1.16c; Willetts & Bullock,<br />

1992) and fruit bodies (Williams et al., 1985).<br />

The composition of proteins on the surface of<br />

hyphal walls may also play an important role in<br />

recognition and adhesion phenomena (de Nobel<br />

et al., 2001).<br />

1.3.1 Mycelial strands<br />

The formation of aggregates of parallel, relatively<br />

undifferentiated hyphae is quite common<br />

in the Basidiomycota and in some Ascomycota.<br />

For instance, mycelial strands form the familiar<br />

‘spawn’ of the cultivated mushroom Agaricus<br />

bisporus. Strands arise most readily from a<br />

well-developed mycelium extending from an<br />

exhausted food base in<strong>to</strong> nutrient-poor surroundings<br />

(Fig. 1.12a). When a strand encounters<br />

a source of nutrients exceeding its internal<br />

supply, coherence is lost and a spreading<br />

assimilative mycelium regrows (Moore, 1994).<br />

Alternatively, mycelial strands may be employed<br />

by fungi which produce their fructifications<br />

some distance away from the food base, as in<br />

the stinkhorn, Phallus impudicus. Here the mycelial<br />

strand is more tightly aggregated and is<br />

referred <strong>to</strong> as a mycelial cord. The tip of the<br />

mycelial cord, which arises from a buried tree<br />

stump, differentiates in<strong>to</strong> an egg-like basidiocarp<br />

initially upon reaching the soil surface<br />

(Fig. 1.12b).<br />

The development of A. bisporus strands has<br />

been described by Mathew (1961). Robust leading<br />

hyphae extend from the food base and branch at<br />

fairly wide intervals <strong>to</strong> form finer laterals, most<br />

of which grow away from the parent hypha.<br />

A few branch hyphae, however, form at an acute<br />

angle <strong>to</strong> the parent hypha and tend <strong>to</strong> grow<br />

parallel <strong>to</strong> it. Hyphae of many fungi occasionally<br />

Fig1.12 Mycelial strands. (a) Strands of Podosordaria tulasnei (Ascomycota) extending from a previously colonized rabbit pellet<br />

(arrow) over sand. Note the dissolution of the strand upon reaching a new nutrient source, in this case fresh sterile rabbit pellets.<br />

(b) Excavated mycelial cords of the stinkhorn Phallus impudicus, which can be traced back from the egg-like basidiocarp primordium<br />

<strong>to</strong> the base of an old tree stump (below the bot<strong>to</strong>m of the picture, not shown).


16 INTRODUCTION<br />

grow alongside each other or another physical<br />

obstacle which they chance <strong>to</strong> encounter. A later<br />

and specific stage in strand development is<br />

characterized by the formation of numerous<br />

fine, aseptate ‘tendril hyphae’ as branches from<br />

the older regions of the main hyphae. The tendril<br />

hyphae, which may extend forwards or backwards,<br />

become appressed <strong>to</strong> the main hypha and<br />

branch frequently <strong>to</strong> form even finer tendrils<br />

which grow round the main hyphae and<br />

ensheath them. Major strands are consolidated<br />

by anas<strong>to</strong>moses between their hyphae, and they<br />

increase in thickness by the assimilation of<br />

minor strands. A similar development has been<br />

noted in the strands of Serpula lacrymans, the dryrot<br />

fungus (Fig. 1.13), which are capable of<br />

extending for several metres across brickwork<br />

and other surfaces from a food base in decaying<br />

wood (Jennings & Watkinson, 1982; Nuss et al.,<br />

1991).<br />

By recovering the nutrients from obsolete<br />

strands and forming new strands, colonies can<br />

move about and explore their vicinity in the<br />

search for new food bases (Cooke & Rayner, 1984;<br />

Boddy, 1993). Mycelial strands are capable of<br />

translocating nutrients and water in both directions<br />

(Boddy, 1993; Jennings, 1995). This property<br />

is important not only for decomposer fungi, but<br />

also for species forming mycorrhizal symbioses<br />

with the roots of plants, many of which produce<br />

hyphal strands (Read, 1991).<br />

Fig1.13 The tip of a hyphal strand of Serpula lacrymans<br />

(Basidiomycota). Note the formation of lateral branches<br />

which grow parallel <strong>to</strong> the direction of the main hyphae.The<br />

buckle-shaped structures at the septa are clamp connections.<br />

1.3.2 Rhizomorphs<br />

In contrast <strong>to</strong> mycelial strands or cords which<br />

consist of relatively undifferentiated aggregations<br />

of hyphae and are produced by a great<br />

variety of fungi, rhizomorphs are found in<br />

only relatively few species and contain highly<br />

differentiated tissues. Well-known examples of<br />

rhizomorph-forming fungi are provided by<br />

Armillaria spp. (Figs. 1.14 and 18.13b), which are<br />

serious parasites of trees and shrubs. In<br />

Armillaria, a central core of larger, thin-walled,<br />

elongated cells embedded in mucilage is<br />

surrounded by a rind of small, thicker-walled<br />

cells which are darkly pigmented due <strong>to</strong> melanin<br />

deposition in their walls. These root-like aggregations<br />

are a means for Armillaria <strong>to</strong> spread<br />

underground from one tree root system <strong>to</strong><br />

another. In nature, two kinds are found a<br />

dark, cylindrical type and a paler, flatter type.<br />

The latter is particularly common beneath the<br />

bark of infected trees (see p. 546). Rhizomorphs<br />

on dead trees measure up <strong>to</strong> 4 mm in diameter. It<br />

has been estimated that a rhizomorph only<br />

1 mm in diameter must contain over 1000<br />

hyphae aggregated <strong>to</strong>gether. The development<br />

of rhizomorphs in agar culture has been<br />

described by Garrett (1953, 1970) and Snider<br />

(1959). Initiation of rhizomorphs can first be


HYPHAL AGGREGATES<br />

17<br />

Fig1.14 Rhizomorph structure of Armillaria<br />

mellea (Basidiomycota). (a) Longitudinal<br />

section. (b) Transverse section, diagrammatic.<br />

(c) L.S. diagrammatic. (d) T.S. showing details of<br />

cells in the rind (r), cortex (c) and medulla (m).<br />

(e)L.S.showingdetailsofcells.<br />

observed after about 7 days’ mycelial growth on<br />

the agar surface as a compact mass of darkly<br />

pigmented hypertrophied cells. These pigmented<br />

structures have been termed microsclerotia.<br />

From white, non-pigmented points on their<br />

surface, the rhizomorphs develop. The growth<br />

of rhizomorphs can be several times faster than<br />

that of unorganized hyphae (Rishbeth, 1968).<br />

The most striking feature of the development of<br />

rhizomorphs is their compact growing point at<br />

the apex, which consists of small isodiametric<br />

cells protected by an apical cap of intertwined<br />

hyphae immersed in mucilage which they<br />

produce. Because of its striking similarity with<br />

a growing plant root, the rhizomorph tip was<br />

initially interpreted as a meristematic zone<br />

(Motta, 1967), but its hyphal nature can be<br />

demonstrated by careful ultrastructural observations<br />

(Powell & Rayner, 1983; Rayner et al., 1985).<br />

Behind the apex there is a zone of elongation.<br />

The centre of the rhizomorph may be hollow or<br />

solid. Surrounding the central lumen or making<br />

up the central medulla is a zone of enlarged<br />

hyphae 4 5 times wider than the vegetative<br />

hyphae (Fig. 1.14e). Possibly these vessel hyphae<br />

serve in translocation (Cairney, 1992; Jennings,<br />

1995). Towards the periphery of the rhizomorph,<br />

the cells become smaller, darker, and thicker<br />

walled. Extending outwards between the outer<br />

cells of the rhizomorph, there may be a growth<br />

of vegetative hyphae somewhat resembling the<br />

root-hair zone in a higher plant. Rhizomorphs<br />

may develop on monokaryotic mycelia derived<br />

from single basidiospores, or on dikaryotic


18 INTRODUCTION<br />

Fig1.15 Rhizomorphs of Podosordariatulasnei (Ascomycota). (a) Subterraneanrhizomorphsby which the fungus spreads through the<br />

soil. (b) T.S. showing the dark rind (1 2 cells thick) and a cortex consisting of thick-walled hyaline cells.<br />

mycelia following fusion of compatible monokaryotic<br />

hyphae. Dikaryotic rhizomorphs of<br />

Armillaria do not possess clamp connections<br />

(Hintikka, 1973).<br />

Rhizomorphs are also produced by other<br />

Basidiomycota and a few Ascomycota (Fig. 1.15;<br />

Webster & Weber, 2000). They are mainly formed<br />

in soil. An interesting exception is presented by<br />

tropical Marasmius spp., which form a network<br />

of aerial rhizomorphs capable of intercepting<br />

falling leaves before they reach the ground<br />

(Hedger et al., 1993). Because these rhizomorphs<br />

have a rudimentary fruit body cap at their<br />

extending apex (Hedger et al., 1993), they have<br />

been interpreted as indefinitely extending fruit<br />

body stipes (Moore, 1994). Mycelial strands and<br />

rhizomorphs represent extremes in a range of<br />

hyphal aggregations, and several intergrading<br />

forms can be recognized (Rayner et al., 1985).<br />

1.3.3 Sclerotia<br />

Sclerotia are pseudoparenchyma<strong>to</strong>us aggregations<br />

of hyphae embedded in an extracellular<br />

glucan matrix. A hard melanized rind may be<br />

present or absent. Sclerotia serve a survival<br />

function and contain intrahyphal s<strong>to</strong>rage<br />

reserves such as polyphosphate, glycogen,<br />

protein, and lipid (Willetts & Bullock, 1992).<br />

The glucan matrix, <strong>to</strong>o, may be utilized as a<br />

carbohydrate source during sclerotium germination<br />

(Backhouse & Willetts, 1985). Sclerotia may<br />

also have a reproductive role and are the only<br />

known means of reproduction in certain species.<br />

They are produced by a relatively small number<br />

of Asco- and Basidiomycota, especially plantpathogenic<br />

species such as Rhizoc<strong>to</strong>nia spp.<br />

(p. 595), Sclerotinia spp. (p. 429) and Claviceps<br />

purpurea (p. 349). The form of sclerotia is very<br />

variable (Butler, 1966). The subterranean sclerotium<br />

of the Australian Polyporus mylittae (see Figs.<br />

18.13c,d) can reach the size of a football and is<br />

known as native bread or blackfellow’s bread. At<br />

the other extreme, they may be of microscopic<br />

dimensions consisting of a few cells only. Several<br />

kinds of development in sclerotia have been<br />

distinguished (Townsend & Willetts, 1954;<br />

Willetts, 1972).<br />

The loose type<br />

This is exemplified by Rhizoc<strong>to</strong>nia spp., which<br />

are sclerotial forms of fungi belonging <strong>to</strong> the<br />

Basidiomycota. Sclerotia of the loose type are<br />

readily seen as the thin brownish-black<br />

scurfy scales so common on the surface of


HYPHAL AGGREGATES<br />

19<br />

Fig1.16 Development of sclerotia. (a) The loose type, as seen in Rhizoc<strong>to</strong>nia (Moniliopsis) solani.(b)HyphaofBotrytiscinerea showing<br />

dicho<strong>to</strong>mous branching on a glass coverslip <strong>to</strong> initiate the terminal type of sclerotium. (c) Later stage of sclerotium formation<br />

in B. cinerea.The hyphae have become melanized and are growing away from the glass surface.They are embedded in a glucan<br />

matrix (arrows). (d) Mature sclerotia of B. cinerea on a stem of Conium. Some sclerotia are germinating <strong>to</strong> produce tufts of<br />

conidiophores. (e) Sclerotia of Claviceps purpurea from an ear of rye (Secale cereale). Rye grains are shown for size comparison.<br />

(a) and (b) <strong>to</strong> same scale.<br />

pota<strong>to</strong> tubers. In pure culture, sclerotial initials<br />

arise by branching and septation of hyphae<br />

(Fig. 1.16a). These cells become filled with dense<br />

contents and numerous vacuoles, and darken <strong>to</strong><br />

reddish-brown. The mature sclerotium does not<br />

show well-defined zones or ‘tissues’. It is made<br />

up of a central part which is pseudoparenchyma<strong>to</strong>us,<br />

although its hyphal nature can be seen.<br />

Towards the outside, the hyphae are more<br />

loosely arranged; a rind of thick-walled hyphae<br />

is absent (Willetts, 1969).<br />

The terminal type<br />

This form is characterized by a well-defined<br />

pattern of branching. It is produced, for example,<br />

by Botrytis cinerea, the cause of grey mould<br />

diseases on a wide range of plants, and by the<br />

saprotrophic Pyronema domesticum (see p. 415).<br />

Sclerotia of B. cinerea are found on overwintering<br />

stems of herbaceous plants, especially<br />

umbellifers such as Angelica, Anthriscus, Conium<br />

and Heracleum. They can also be induced <strong>to</strong> form<br />

in culture, especially on agar media with a high


20 INTRODUCTION<br />

carbon/nitrogen ratio. When growing on host<br />

tissue, the sclerotia of Botrytis may include host<br />

cells, a feature shared also by sclerotia of<br />

Sclerotinia spp. <strong>to</strong> which Botrytis is related (see<br />

p. 429). Sclerotia arise by repeated dicho<strong>to</strong>mous<br />

branching of hyphae, accompanied by cross-wall<br />

formation (Fig. 1.16b). The hyphae then aggregate,<br />

melanize and produce mucilage, giving the<br />

appearance of a solid tissue (Fig. 1.16c). A mature<br />

sclerotium may be about 10 mm long and<br />

3 5 mm wide, and is usually flattened, measuring<br />

1 3 mm in thickness. It is often orientated<br />

parallel <strong>to</strong> the long axis of the host plant<br />

(Fig. 1.16d). It is differentiated in<strong>to</strong> a rind<br />

composed of several layers of rounded, dark<br />

cells, a narrow cortex of thin-walled pseudoparenchyma<strong>to</strong>us<br />

cells with dense contents, and a<br />

medulla made up of loosely arranged filaments.<br />

Nutrient reserves are s<strong>to</strong>red in the cortical and<br />

medullary regions (Willetts & Bullock, 1992).<br />

The strand type<br />

Sclerotinia gladioli, the causal agent of dry rot of<br />

corms of Gladiolus, Crocus and other plants,<br />

forms sclerotia of this type. Sclerotial initials<br />

commence with the formation of numerous<br />

side branches which arise from one or more<br />

main hyphae. Where several hyphae are<br />

involved, they lie parallel. They are thicker than<br />

normal vegetative hyphae, and become divided<br />

by septa in<strong>to</strong> chains of short cells. These cells<br />

may give rise <strong>to</strong> short branches, some of which<br />

lie parallel <strong>to</strong> the parent hypha, whilst others<br />

grow out at right angles and branch again before<br />

coalescing. The hyphae at the margin continue<br />

<strong>to</strong> branch, and the whole structure darkens.<br />

The mature sclerotium is about 0.1 0.3 mm in<br />

diameter, and is differentiated in<strong>to</strong> a rind of<br />

small, thick-walled cells and a medulla of large,<br />

thin-walled hyphae. More complex sclerotia are<br />

found in Sclerotium rolfsii, the sclerotial state of<br />

Pellicularia rolfsii (Basidiomycota). Here the<br />

mature sclerotium is differentiated in<strong>to</strong> four<br />

zones: a fairly thick skin or cuticle, a rind made<br />

up of 2 4 layers of tangentially flattened cells, a<br />

cortex of thin-walled cells with densely staining<br />

contents, and a medulla of loose filamen<strong>to</strong>us<br />

hyphae with dense contents. Chet et al. (1969)<br />

have shown that the skin or cuticle is made up of<br />

the remnants of cell walls attached <strong>to</strong> the<br />

outside of the empty, melanized, thick-walled<br />

rind cells. All the cells of the strand-type<br />

sclerotium have thicker walls than those of<br />

vegetative hyphae. Cells of the outer cortex<br />

contain large s<strong>to</strong>rage bodies which consist of<br />

protein (Kohn & Grenville, 1989) and leave little<br />

room for cy<strong>to</strong>plasm or other organelles. The<br />

inner cortex is also densely packed with s<strong>to</strong>rage<br />

granules.<br />

Other types<br />

There is a great diversity of other types of<br />

sclerotia (Butler, 1966). The sclerotia of Claviceps<br />

purpurea, the ‘ergots’ of grasses and cereals<br />

(Fig. 1.16e; see also p. 349), develop from a preexisting<br />

mass of mycelium which fills and<br />

replaces the cereal ovary, starting from the base<br />

and extending <strong>to</strong>wards the apex. The outer layers<br />

form a violet, dark grey or black rind enclosing<br />

colourless, thick-walled cells. These contain<br />

abundant s<strong>to</strong>rage lipids which constitute 45%<br />

of the dry weight of a C. purpurea sclerotium<br />

(Kybal, 1964). Cordyceps militaris, an insect parasite,<br />

forms a dense mass of mycelium in the<br />

buried insect’s body (p. 360). This mass of<br />

mycelium, from which fructifications develop,<br />

is enclosed by the exoskele<strong>to</strong>n of the host, not by<br />

a fungal rind. Many wood-rotting fungi enclose<br />

colonized woody tissue with a black zone-line of<br />

dark, thick-walled cells, and the whole structure<br />

may be regarded as a kind of sclerotium.<br />

The giant sclerotium of Polyporus mylittae is<br />

marbled in structure, comprising white strata<br />

and translucent tissue. It has an outer, smooth,<br />

thin black rind. Three distinct types of hyphae<br />

make up the tissues: thin-walled, thick-walled<br />

and ‘layered’ hyphae. Thin- and thick-walled<br />

hyphae are abundant in the white strata but<br />

sparse in the translucent tissue, whereas the<br />

layered hyphae occur only in the translucent<br />

tissue. Detached sclerotia are capable of forming<br />

basidiocarps without wetting. It is believed that<br />

the translucent tissue functions as an extracellular<br />

nutrient and water s<strong>to</strong>re (Macfarlane et al.,<br />

1978). The structure of the sclerotium appears <strong>to</strong><br />

be related <strong>to</strong> its ability <strong>to</strong> fruit in dry conditions,<br />

such as occur in Western Australia.


HYPHAL AGGREGATES<br />

21<br />

Germination of sclerotia<br />

Sclerotia can survive for long periods, sometimes<br />

for several years (Coley-Smith & Cooke, 1971;<br />

Willetts, 1971). Germination may take place in<br />

three ways by the development of mycelium,<br />

asexual spores (conidia) or sexual fruit bodies<br />

(ascocarps or basidiocarps). Mycelial germination<br />

occurs in Sclerotium cepivorum, the cause<br />

of white-rot of onion, and is stimulated by<br />

volatile exudates from onion roots (see p. 434).<br />

Conidial development occurs in Botrytis cinerea<br />

and can be demonstrated by placing overwintered<br />

sclerotia in moist warm conditions<br />

(Fig. 1.16d; Weber & Webster, 2003). The development<br />

of ascocarps (i.e. carpogenic germination)<br />

is seen in Sclerotinia, where stalked cups<br />

or apothecia, bearing asci, arise from sclerotia<br />

under suitable conditions (Fig. 15.2), and in<br />

Claviceps purpurea, where the overwintered sclerotia<br />

give rise <strong>to</strong> a perithecial stroma (Fig. 12.26c).<br />

Depending on environmental conditions, the<br />

sclerotia of some species may respond by<br />

germinating in different ways.<br />

1.3.4 The mantle of ec<strong>to</strong>mycorrhiza<br />

The root tips of many coniferous and deciduous<br />

trees with ec<strong>to</strong>mycorrhizal associations,<br />

especially those growing in relatively infertile<br />

soils, are covered by a mantle. This is a<br />

continuous sheet of fungal hyphae, several<br />

layers thick (see Fig. 19.10). The mycelium<br />

extends outwards in<strong>to</strong> the litter layer of the<br />

soil, and inwards as single hyphae growing<br />

intercellularly, i.e. between the outer cortical<br />

cells of the root, <strong>to</strong> form the so-called Hartig<br />

net. Hyphae growing outwards from the<br />

mantle effectively replace the root hairs as a<br />

system for the absorption of minerals from the<br />

soil, and there is good evidence that, in most<br />

normal forest soils of low <strong>to</strong> moderate fertility,<br />

the performance and nutrient status<br />

of mycorrhizal trees is superior <strong>to</strong> that of<br />

uninfected trees (Smith & Read, 1997).<br />

Most fungi causing ec<strong>to</strong>mycorrhizal infections<br />

are Basidiomycota, especially members of the<br />

Homobasidiomycetes (pp. 526 and 581). Within<br />

the soil or in pure culture, mycelial<br />

strands may form, but the mycelium is not<br />

aggregated in<strong>to</strong> the tissue-like structure of the<br />

mantle.<br />

1.3.5 Fruit bodies of Ascomycota and<br />

imperfect fungi<br />

In the higher fungi, hyphae may aggregate in a<br />

highly regulated fashion <strong>to</strong> form fruiting structures<br />

which are an important and often speciesspecific<br />

feature of identification. In the<br />

Ascomycota, the fruit bodies produce sexual<br />

spores (i.e. as the result of nuclear fusion and<br />

meiosis) which are termed ascospores and are<br />

contained in globose or cylindrical cells called<br />

asci (Lat. ascus ¼ a sac, tube). In most cases,<br />

the asci can discharge their ascospores explosively.<br />

Asci, although occasionally naked, are<br />

usually enclosed in an aggregation of hyphae<br />

termed an ascocarp or ascoma. Ascocarps are<br />

very variable in form, and several types have<br />

been distinguished (see Fig. 8.16). Their features<br />

and development will be described more fully<br />

later. Forms in which the asci are <strong>to</strong>tally<br />

enclosed, and in which the ascocarp has no<br />

special opening, are termed cleis<strong>to</strong>thecia. In<br />

contrast, gymnothecia consist of a loose mesh<br />

of hyphae. Both are found in the Plec<strong>to</strong>mycetes<br />

(Chapter 11). A modified cleis<strong>to</strong>thecium is characteristic<br />

of the Erysiphales (Chapter 13). Cup<br />

fungi (Discomycetes, Chapters 14 and 15) possess<br />

saucer-shaped ascocarps termed apothecia, with<br />

a mass of non-fertile hyphae supporting a layer<br />

of asci lining the upper side of the fruit body. The<br />

non-fertile elements of the apothecium often<br />

show considerable differentiation of structure.<br />

The asci in apothecia are free <strong>to</strong> discharge their<br />

ascospores at the same time. In other<br />

Ascomycota, the asci are contained within<br />

ascocarps with a very narrow opening or ostiole,<br />

through which each ascus must discharge its<br />

spores separately. Ascocarps of this type are<br />

termed perithecia or pseudothecia. Perithecia<br />

are found in the Pyrenomycetes (Chapter 12)<br />

whilst pseudothecia occur in the<br />

Loculoascomycetes (Chapter 17). These two<br />

types of ascocarp develop in different ways. In<br />

many of the Pyrenomycetes, the perithecia are<br />

borne on or embedded in a mass of fungal tissue<br />

termed the perithecial stroma, and these are


22 INTRODUCTION<br />

well shown by the Xylariales (p. 332), and by<br />

Cordyceps (p. 360) and Claviceps (p. 349). In some<br />

cases, in addition <strong>to</strong> the perithecial stroma, a<br />

fungus may develop a stromatic tissue on or<br />

within which asexual spores (conidia) develop.<br />

Nectria cinnabarina (p. 341), the coral spot fungus<br />

so common on freshly dead deciduous twigs, is<br />

such an example. It initially forms pink conidial<br />

stromata which later, under suitable conditions<br />

of humidity, become converted in<strong>to</strong> perithecial<br />

stromata.<br />

Among the imperfect (asexual) fungi, mycelial<br />

aggregations bearing conidia are seen in<br />

various genera. In some, there are tufts<br />

of parallel conidiophores termed coremia<br />

or synnemata, exemplified by Penicillium<br />

claviforme (see Fig. 11.19). In some imperfect<br />

fungi formerly called Coelomycetes, the conidia<br />

develop in flask-shaped cavities termed pycnidia<br />

(see Figs. 17.3 17.5). Various other<br />

kinds of mycelial fruiting aggregates are also<br />

known.<br />

1.3.6 Fruit bodies of Basidiomycota<br />

The fruit bodies of mushrooms, <strong>to</strong>ads<strong>to</strong>ols,<br />

bracket fungi, etc., are all examples of basidiocarps<br />

or basidiomata which bear the sexually<br />

produced spores (basidiospores) on basidia.<br />

Basidiocarps are almost invariably constructed<br />

from dikaryotic hyphae, but how vegetative<br />

hyphae aggregate <strong>to</strong> form a mushroom fruit<br />

body is still a mystery (Moore, 1994). Wessels<br />

(1997) has suggested that hydrophobins coating<br />

the surface of hyphae may confer adhesive<br />

properties, leading <strong>to</strong> their aggregation <strong>to</strong> form<br />

a fruit body initial as the first step in morphogenesis.<br />

Once an initial has been formed, its<br />

glucan matrix may provide an environment for<br />

the exchange of signalling molecules between<br />

hyphae. Moore (1994) speculated that morphogenesis<br />

might ultimately be determined by<br />

induction hyphae exerting a control over<br />

surrounding hyphae, leading <strong>to</strong> the development<br />

of morphogenetic units. This morphogenetic<br />

commitment must happen at a very early stage.<br />

For instance, in the ink-cap (Coprinus cinereus) an<br />

initial measuring only 1% of the final fruit body<br />

size is already differentiated in<strong>to</strong> stipe and cap<br />

(Moore et al., 1979). Therefore, when a mushroom<br />

fruit body expands, this is due mainly <strong>to</strong> the<br />

enlargement of existing hyphae, whereas new<br />

apical growth is restricted mainly <strong>to</strong> branches<br />

filling up the space generated during expansion<br />

(Moore, 1994). Hyphae making up the mature<br />

basidiocarp may show considerable differentiation<br />

in structure and function. This is perhaps<br />

most highly developed in polypore-type basidiocarps,<br />

where a number of morphologically<br />

distinct hyphal types have been recognized<br />

(p. 517).<br />

1.4 Spores of fungi<br />

The reproduction by means of small spores is<br />

a corners<strong>to</strong>ne in the ecology of fungi. Although<br />

a single spore may have a negligible chance of<br />

reaching a suitable substrate, spores may be<br />

produced in such quantities that even discrete<br />

substrates can be exploited by the species as a<br />

whole. Only a few fungi make do without spores,<br />

surviving solely by means of mycelium and<br />

sclerotia. Spores may be organs of sexual or<br />

asexual reproduction, and they are involved in<br />

dispersal and survival. Gregory (1966) distinguished<br />

between xenospores (Gr. xenos ¼<br />

a foreigner) for spores which are dispersed from<br />

their place of origin and memnospores<br />

(Gr. mémnon ¼ steadfast, <strong>to</strong> persist), which stay<br />

where they were formed. Some spores are<br />

violently discharged from the organs which<br />

bear them, energy for dispersal being provided<br />

by the spore itself or the structure producing<br />

it (Ingold, 1971). However, many spores are<br />

dispersed passively by the action of gravity, air<br />

or water currents, rain splash, or by animals,<br />

especially insects. Dispersal may also occur by<br />

human traffic. Spores may be present in the<br />

outdoor air at such high concentrations (e.g.<br />

100 Cladosporium spores l 1 ) that they can cause<br />

allergic respira<strong>to</strong>ry diseases when inhaled (Lacey,<br />

1996). In freshwater, the asexually produced<br />

spores (conidia) of aquatic hyphomycetes, which<br />

colonize autumn-shed tree leaves, may reach<br />

concentrations of 10 000 20 000 spores l 1 (see<br />

p. 685). Long-range dispersal of air-borne spores


SPORES OF FUNGI<br />

23<br />

over thousands of kilometres is known <strong>to</strong> occur<br />

in nature. For instance, the urediniospores of<br />

the coffee rust fungus, Hemileia vastatrix, are<br />

thought <strong>to</strong> have travelled from Africa <strong>to</strong> South<br />

America by wind at high altitudes, and the<br />

urediniospores of black stem rust of wheat<br />

(Puccinia graminis) undergo an annual migration<br />

from states bordering the Gulf of Mexico <strong>to</strong><br />

the prairies of North America and Canada<br />

(Fig. 22.11). These spores are protected from<br />

the deleterious effects of UV irradiation in<br />

the upper atmosphere by pigments in the<br />

spore wall.<br />

Some spores are not dispersed but survive<br />

in situ, e.g. the oospores of many soil-inhabiting<br />

Oomycota (Chapter 5), the zygospores of<br />

Zygomycota (Chapter 7) and the chlamydospores<br />

of Glomales (see p. 217) and other fungi. Fungal<br />

spores may remain dormant for many years,<br />

especially under dry and cold conditions<br />

(Sussman & Halvorson, 1966; Sussman, 1968).<br />

An extreme example of spore survival is shown<br />

by the recovery of viable spores of several fungi<br />

from glacial ice cores, including those of<br />

Cladosporium cladosporioides from ice samples<br />

4500 years old (Ma et al., 2000).<br />

The morphology and structure of fungal<br />

spores show great variability, from unicellular<br />

<strong>to</strong> multicellular, branched or unbranched or<br />

sometimes spirally coiled, thin- or thick-walled<br />

with hyaline or pigmented walls, dry or sticky,<br />

smooth or ornamented by mucilaginous extensions,<br />

spines, folds or reticulations. A number of<br />

general descriptive terms have been applied <strong>to</strong><br />

characterize spores in relation <strong>to</strong> the number of<br />

cells and septa which they contain. Single-celled<br />

spores are termed amerospores (Gr. a ¼ not,<br />

meros ¼ a part; i.e. not divided), two-celled spores<br />

are didymospores (Gr. didymos ¼ double), spores<br />

with more than one transverse septum are<br />

phragmospores (Gr. phragmos ¼ a hedge, barricade),<br />

and spores with transverse and longitudinal<br />

septa are dictyospores (Gr. dictyon ¼<br />

a net). These terms may be qualified by prefixes<br />

indicating spore pigmentation such as hyalo- for<br />

colourless (hyaline) spores and phaeo- for spores<br />

with dark-coloured (melanized) walls.<br />

Special terms have also been used <strong>to</strong> refer <strong>to</strong><br />

spore shape. Scolecospores (Gr. skolex ¼ a worm)<br />

are worm-shaped, helicospores (Gr. helix ¼<br />

twisted or wound) are spores with a two- or<br />

three-dimensional spiral shape, whilst staurospores<br />

(Gr. stauros ¼ a cross) have arms radiating<br />

from a central point or axis. Spore septation,<br />

colour and shape, along with other criteria such<br />

as the arrangement of structures which bear the<br />

spores, have been used in classification and<br />

identification, especially in conidial fungi<br />

which do not show sexual reproduction. These<br />

criteria rarely lead <strong>to</strong> natural systems of classification,<br />

but <strong>to</strong> ‘form genera’ or ‘anamorph<br />

genera’ made up of species unified by having<br />

similar spore forms.<br />

Some of the more common spore types<br />

are described below. There are numerous<br />

other, less-common kinds of spore found in<br />

fungi, and they are described later, in relation<br />

<strong>to</strong> the particular fungal groups in which they<br />

occur.<br />

1.4.1 Zoospores<br />

These are spores which are self-propelled by<br />

means of flagella. Propulsion is often coupled<br />

with chemotactic movement, zoospores having<br />

the ability <strong>to</strong> sense chemicals diffusing from<br />

suitable substrata and <strong>to</strong> move <strong>to</strong>wards them, or<br />

gametes detecting and following extremely low<br />

concentrations of hormones. In some cases<br />

oxygen or light are also stimuli for tactic movement.<br />

The fungal groups which possess flagella<br />

are mostly aquatic or, if terrestrial, rely on water<br />

for dispersal or infection. Their zoospores are of<br />

four kinds (see Fig. 1.17):<br />

1. Posteriorly flagellate zoospores with<br />

flagella of the whiplash type are characteristic<br />

of the Chytridiomycota (Chapter 6). Each<br />

whiplash flagellum has 11 microtubules<br />

arranged in the 9 þ 2 pattern typical of<br />

eukaryotes. The microtubules are enclosed in<br />

a smooth, membranous axoneme sheath continuous<br />

with the plasma membrane. In most<br />

members of the Chytridiomycota there is a<br />

single posterior flagellum (Fig. 1.17a), but in<br />

the rumen-inhabiting Neocallimastigales there<br />

may be up <strong>to</strong> 16 flagella (Fig. 1.17b). Such spores<br />

are driven forward by sinusoidal rhythmic<br />

beating of the flagellum. This type of zoospore


24 INTRODUCTION<br />

Fig1.17 Zoospore types found in fungi, diagrammatic and not <strong>to</strong> scale.The arrow indicates the direction of movement of the<br />

zoospore. (a) Posteriorly uniflagellate (opisthokont) zoospore with a flagellum of the whiplash type found in many Chytridiomycota.<br />

(b) Posteriorly multiflagellate zoospore with numerous (up <strong>to</strong>16) whiplash flagella which occur in certain anaerobic rumen-inhabiting<br />

Chytridiomycota (Neocallimastigales). (c) Zoospore with unequal (anisokont) whiplash flagella characteristic of the Myxomycota<br />

and the Plasmodiophoromycota. (d) Anteriorly uniflagellate zoospore with a flagellum of the tinsel type, the axoneme being clothed<br />

with rows of mastigonemes, typical of the Hyphochytriomycota. (e,f) Biflagellate zoospores with heterokont flagella, one of the<br />

whiplash and the other of the tinsel type, which are found in different groups of the Oomycota. For more details turn <strong>to</strong> the<br />

different fungal groups.<br />

flagellation is termed opisthokont (Gr. opisthen ¼<br />

behind, at the back; kon<strong>to</strong>s ¼ a pole). Detailed<br />

descriptions of the fine structure of chytridiomycete<br />

zoospores are given on p. 129.<br />

2. Biflagellate zoospores with two whiplash<br />

flagella of unequal length are called anisokont<br />

(Fig. 1.17c) and are found in some Myxomycota<br />

and the Plasmodiophoromycota, both now<br />

classified among the Pro<strong>to</strong>zoa (see Chapters 2<br />

and 3).<br />

3. Anteriorly uniflagellate zoospores with<br />

a flagellum of the tinsel type are characteristic<br />

of the Hyphochytriomycota (Chapter 4). The<br />

axoneme sheath of the tinsel or straminipilous<br />

flagellum (Lat. stramen ¼ straw; pilus ¼ hair) is<br />

adorned by two rows of fine hairs (Fig. 1.17d).<br />

These are called tripartite tubular hairs or<br />

mastigonemes (Gr. mastigion ¼ a small whip;<br />

nema ¼ a thread). Rhythmic sinusoidal beating of<br />

the tinsel type flagellum pulls the zoospore<br />

along, in contrast <strong>to</strong> the pushing action of<br />

whiplash flagellum. Details of the fine structure<br />

of this type of zoospore are given in Fig. 4.5.<br />

4. Biflagellate zoospores with anteriorly or<br />

laterally attached flagella, one of which is of<br />

the whiplash type and the other of the tinsel<br />

type (Figs. 1.17e,f), are characteristic of the<br />

Oomycota (Chapter 5). Zoospores with the two<br />

different kinds of flagellum are heterokont.<br />

Where the two types of flagellum are attached<br />

anteriorly, as in the first-released zoospores of<br />

Saprolegnia, their propulsive actions tend <strong>to</strong><br />

work against each other and the zoospore is<br />

a very poor swimmer (Fig. 1.17e). However,<br />

the secondary zoospore (termed the principal<br />

zoospore) in Saprolegnia and in many other<br />

Oomycota has laterally attached flagella, with<br />

the tinsel-type (pulling action) flagellum pointing<br />

forwards and the whiplash-type (pushing<br />

action) flagellum directed backwards and possibly<br />

acting as a rudder, jointly providing much<br />

more effective propulsion (Fig. 1.17f).<br />

1.4.2 Sporangiospores<br />

In the Zygomycota, and especially in the<br />

Mucorales (see p. 180), the asexual spores are<br />

contained in globose sporangia (Fig. 1.18) or<br />

cylindrical merosporangia. Because they are<br />

non-motile, the spores are sometimes termed<br />

aplanospores (Gr. a ¼ not, planos ¼ roaming).


SPORES OF FUNGI<br />

25<br />

Fig1.18 Sporangia in Mortierella (Umbelopsis)<br />

vinacea. (a) Maturing sporangium in which<br />

the cy<strong>to</strong>plasm is being cleaved in<strong>to</strong><br />

numerous sporangiospores. (b) Release of<br />

sporangiospores by breakdown of the<br />

sporangial wall. Unusually, in M. vinacea the<br />

sporangiospores are angular in shape.<br />

The spores may be uni- or multinucleate and are<br />

unicellular. They generally have thin, smooth<br />

walls and are almost always globose or ellipsoid<br />

in shape. They are formed by cleavage of the<br />

sporangial cy<strong>to</strong>plasm. They vary in colour from<br />

hyaline (colourless) <strong>to</strong> yellow, due <strong>to</strong> carotenoid<br />

pigments in the cy<strong>to</strong>plasm. When mature, they<br />

may be surrounded by mucilage, in which case<br />

they are usually dispersed by rain splash or<br />

insects, or they may be dry and dispersed by wind<br />

currents. In some genera, e.g. Pilobolus, entire<br />

sporangia become detached. The number of<br />

sporangiospores per sporangium may vary from<br />

several thousand <strong>to</strong> only one. The detachment<br />

and dispersal of intact sporangia containing a<br />

few sporangiospores or a single one is indicative<br />

of the way in which conidia may have evolved<br />

from one-spored sporangia.<br />

1.4.3 Ascospores<br />

Ascospores are the characteristic spores of the<br />

largest group of fungi, the Ascomycota or<br />

ascomycetes. They are meiospores and are<br />

formed in the developing ascus as a result of<br />

nuclear fusion immediately followed by meiosis.<br />

The four haploid daughter nuclei then divide<br />

mi<strong>to</strong>tically <strong>to</strong> give eight haploid nuclei around<br />

which the ascospores are cut out. Details of<br />

ascospore development are described in Fig. 8.11.<br />

In most ascomycetes, the eight ascospores are<br />

contained within a cylindrical ascus, from which<br />

they are squirted out <strong>to</strong>gether with the ascus<br />

sap when the tip of the turgid ascus breaks down<br />

and the elastic ascus walls contract. The distance<br />

of discharge may be 1 cm or more. In some cases,<br />

for example, the Plec<strong>to</strong>mycetes (Chapter 11)<br />

and in ascomycetes with subterranean fruit<br />

bodies, such as the false truffles (Elaphomyces<br />

spp.; Fig. 11.21) and truffles (Tuber spp. and their<br />

allies; p. 423), ascospore release is non-violent<br />

and their asci are not cylindrical but globose.<br />

Ascospores vary greatly in size, shape and colour.<br />

In size, the range is from about 4 5 1 mm in<br />

small-spored forms such as the minute cup<br />

fungus Dasyscyphus, <strong>to</strong> 130 45 mm in the lichen<br />

Pertusaria pertusa. The shape of ascospores varies<br />

from globose <strong>to</strong> oval, elliptical, lemon-shaped,<br />

sausage-shaped, cylindrical, or needle-shaped.<br />

Ascospores are often asymmetric in form with<br />

a wider, blunter, anterior part and a narrower,<br />

more tapering posterior. This shape increases<br />

their acceleration as they are squeezed out<br />

through the opening of the ascus. Ascospores<br />

may be uninucleate or multinucleate, unicellular<br />

or multicellular, divided up by transverse or<br />

by transverse and longitudinal septa. In some


26 INTRODUCTION<br />

genera, e.g. Hypocrea (Fig. 12.15) or Cordyceps<br />

(Fig. 12.33), the multicellular ascospores may<br />

break up in<strong>to</strong> part-spores within the ascus prior<br />

<strong>to</strong> discharge. The ascospore wall may be thin or<br />

thick, hyaline or coloured, smooth or rough,<br />

sometimes cast in<strong>to</strong> reticulate folds or ornamented<br />

by ridges, and it may have a mucilaginous<br />

outer layer which is sometimes extended <strong>to</strong> form<br />

simple or branched appendages, especially in<br />

marine ascomycetes where they aid buoyancy<br />

and attachment. In many cases, ascospores are<br />

resting structures which survive adverse conditions.<br />

They may have extensive food reserves in<br />

the form of lipids and sugars such as trehalose.<br />

Because the formation of ascospores involves<br />

meiosis, they are important not only as a means<br />

of dispersal and survival but also in genetic<br />

recombination.<br />

It is obvious that there is no such thing as<br />

a typical ascospore. Neurospora tetrasperma will<br />

serve as an example of an ascospore whose<br />

structure has been extensively studied (Lowry &<br />

Sussman, 1958, 1968). This fungus is somewhat<br />

unusual in that it has four-spored asci and the<br />

ascospores are binucleate. The spores are black,<br />

thick-walled and shaped rather like a rugby<br />

football, but with flattened ends. The name<br />

Neurospora refers <strong>to</strong> the ribbed spores, because<br />

the dark outer wall is made up of longitudinal<br />

raised ribs, separated by interrupted grooves. The<br />

structure of a spore in section is shown in<br />

Fig. 1.19. Within the cy<strong>to</strong>plasm of the spore<br />

are the two nuclei, fragments of endoplasmic<br />

reticulum (not illustrated), swollen mi<strong>to</strong>chondria<br />

and vacuoles, bounded by single unit membranes.<br />

The wall surrounding the pro<strong>to</strong>plast is<br />

composed of several layers. The innermost layer<br />

is the endospore, outside of which is the<br />

epispore. The ribbed layer is termed the perispore.<br />

Between the ribs are lighter intercostal<br />

veins containing a material which is chemically<br />

distinct from the ribs. This material is continuous<br />

over the whole surface of the spore, giving<br />

it a relatively smooth surface. The spore germinates<br />

by the extrusion of germ tubes from a preexisting<br />

germ pore, a thin area in the epispore at<br />

either end of the spore. In many ascomycetes a<br />

trigger is required for germination, e.g. heat<br />

shock in Neurospora or a chemical stimulus, for<br />

example in ascomycetes which grow and fruit on<br />

the dung of herbivorous mammals and whose<br />

spores are subjected <strong>to</strong> digestive treatment.<br />

1.4.4 Basidiospores<br />

Basidiospores are the sexual spores which<br />

characterize a large group of fungi, the<br />

Basidiomycota or basidiomycetes. In comparison<br />

with the morphological diversity of ascospores,<br />

basidiospores are more uniform. They also show<br />

a smaller size range, from about 3 <strong>to</strong> 20 mm,<br />

which is possibly related <strong>to</strong> their unique method<br />

of discharge. They are normally found in<br />

groups of four attached by tapering sterigmata<br />

<strong>to</strong> the cell which bears them, the basidium. At<br />

the time of their discharge all basidiospores<br />

Fig1.19 Neurospora tetrasperma.T.S.<br />

ascospore. Simplified diagram based on<br />

an electron micrograph by Lowry in<br />

Sussman and Halvorson (1966).


SPORES OF FUNGI<br />

27<br />

are unicellular, but they may become septate<br />

after release in some members of the<br />

Heterobasidiomycetes (Chapter 21). In shape,<br />

basidiospores are asymmetric and vary from<br />

sub-globose, sausage-shaped, fusoid, <strong>to</strong> almondshaped<br />

(i.e. flattened), and the wall may be<br />

smooth or ornamented with spines, ridges or<br />

folds. The colour of basidiospores is important<br />

for identification. They may be colourless,<br />

white, cream, yellowish, brown, pink, purple<br />

or black. The spore colour may be due <strong>to</strong> pigments<br />

in the spore cy<strong>to</strong>plasm or in the spore<br />

wall. The appearance of pigments in the wall<br />

occurs relatively late in spore development.<br />

This explains the change of colour of the gill<br />

of a domestic mushroom (Agaricus) from pink,<br />

due <strong>to</strong> cy<strong>to</strong>plasmic spore pigments, <strong>to</strong> dark<br />

purplish-brown when mature, due <strong>to</strong> wall<br />

pigments.<br />

The generalized structure of a basidiospore is<br />

illustrated in Fig. 1.20. Most basidiospores have<br />

a flatter adaxial face and a more curved abaxial<br />

face. The point of attachment of the spore <strong>to</strong> the<br />

sterigma is the hilum, which persists as a scar<br />

at the base of a discharged spore. Close <strong>to</strong> the<br />

hilum is a small projection, the hilar appendix.<br />

This is involved in the unique mechanism of<br />

basidiospore discharge, in which a drop of liquid<br />

perched on the hilar appendix coalesces with<br />

a second blob of liquid on the spore surface,<br />

Fig1.20 Generalized view of a median vertical section<br />

through a basidiospore as seen by transmission electron<br />

microscopy.For clarity, structures such as endoplasmic<br />

reticulum and ribosomes are not illustrated. Diagram<br />

based on Agrocybe acericola, after Ruch and Nurtjahja<br />

(1996).


28 INTRODUCTION<br />

creating a momentum which leads <strong>to</strong> acceleration<br />

of the spore (Money, 1998; see p. 493). The<br />

spore is projected for a short distance (usually<br />

less than 2 mm) from the basidium. Violently<br />

projected spores are termed ballis<strong>to</strong>spores (Lat.<br />

ballista ¼ a military engine for throwing large<br />

s<strong>to</strong>nes), but whilst most basidiospores are ballis<strong>to</strong>spores,<br />

some are not. For example, in the<br />

Gasteromycetes (Chapter 20), which include puffballs,<br />

stinkhorns and their allies, violent spore<br />

projection has been lost in the course of<br />

evolution from ances<strong>to</strong>rs which possessed it.<br />

Likewise, the basidiospores of smut fungi<br />

(Ustilaginales, Chapter 23) are not violently<br />

discharged. The term statismospore (Lat. statio<br />

¼ standing still) is sometimes used for a spore<br />

which is not forcibly discharged.<br />

The cy<strong>to</strong>plasm of basidiospores usually<br />

contains a single haploid nucleus resulting<br />

from meiotic division in the basidium; sometimes<br />

a post-meiotic division gives rise <strong>to</strong> two<br />

genetically identical nuclei. The structure of the<br />

wall is complex. In Agrocybe acericola there are<br />

two layers, a thicker, dark-pigmented, electrondense<br />

outer layer or epispore, and a thinner,<br />

electron-transparent inner layer, the endospore<br />

(Ruch & Nurtjahja, 1996; see Fig. 1.20). The<br />

cultivated mushroom, Agaricus bisporus, has a<br />

three-layered wall making up some 35% of the<br />

dry weight of the spore (Rast & Hollenstein,<br />

1977), whereas the wall of the Coprinus cinereus<br />

basidiospore comprises six distinct layers<br />

(McLaughlin, 1977). A his<strong>to</strong>chemical feature of<br />

the walls of some basidiospores is that they<br />

are amyloid, i.e. they include starch-like<br />

material which stains bluish-purple with iodinecontaining<br />

stains such as Melzer’s reagent. This<br />

reaction is used as a taxonomic character.<br />

The amyloid reaction is due <strong>to</strong> the presence of<br />

unbranched, short-chain amylose molecules. It<br />

has been suggested that this ‘fungal starch’ may<br />

aid dormancy by creating a permeability barrier<br />

<strong>to</strong> oxygen in dry spores. When the amyloid<br />

material is dissolved as water becomes available,<br />

dormancy is lost and spore germination can<br />

proceed (Dodd & McCracken, 1972). In some<br />

basidiospores, e.g. those of Coprinus cinereus and<br />

Agrocybe acericola, the basidiospore has a distinct<br />

germ pore at the end opposite <strong>to</strong> the hilum<br />

(see Fig. 1.20). In other basidiomycetes, e.g.<br />

Oudemansiella mucida, Schizophyllum commune and<br />

Flammulina velutipes, the basidiospores have no<br />

specialized pore.<br />

The reserve contents of the spore may vary.<br />

In some species, lipid is the major s<strong>to</strong>rage<br />

product, and there is an apparent lack of<br />

insoluble polysaccharides such as glycogen<br />

(Ruch & Motta, 1987). In other spores, glycogen<br />

predominates. Where lipid is present, germination<br />

may be fuelled by its breakdown and<br />

utilization, but where it is absent spores are<br />

dependent on external nutrient supplies before<br />

germination and further development is possible.<br />

In addition <strong>to</strong> the usual organelle complement,<br />

microbodies are also prominent in<br />

basidiospores. These are single membranebound<br />

organelles often associated with mi<strong>to</strong>chondria<br />

and lipid globules; they may function<br />

as glyoxisomes containing enzymes involved<br />

in the oxidation of lipids (Ruch & Nurtjahja,<br />

1996).<br />

1.4.5 Zygospores<br />

Zygospores are sexually produced resting structures<br />

formed as a result of plasmogamy between<br />

gametangia which are usually equal in size<br />

(Fig. 1.21a). Nuclear fusion may occur early, or<br />

may be delayed until shortly before meiosis and<br />

zygospore germination. Zygospores are typical<br />

of Zygomycota (Chapter 7). They are often large,<br />

thick-walled, warty structures with abundant<br />

lipid reserves and are unsuitable for longdistance<br />

dispersal, usually remaining in the<br />

position in which they were formed and awaiting<br />

suitable conditions for further development.<br />

The gametangia which fuse <strong>to</strong> form the zygospore<br />

may be uninucleate or multinucleate,<br />

and correspondingly the zygospore may have<br />

one, two or many nuclei within it. Zygospore<br />

germination may be by a germ tube or by the<br />

formation of a germ sporangium.<br />

1.4.6 Oospores<br />

An oospore is a sexually produced spore which<br />

develops from unequal gametangial copulation<br />

or markedly unequal (oogamous) gametic fusion<br />

(Fig. 1.21b). It is the characteristic sexually


SPORES OF FUNGI<br />

29<br />

Fig1.21 Sexual resting structures. (a) Zygospore of Rhizopus sexualis.The zygote has been produced by fusion of two gametangia<br />

and has laid down a thick wall with warty ornamentations. (b) Oospore of Phy<strong>to</strong>phthora erythroseptica.The oogonium (o) has grown<br />

through the antheridium (a), and the oosphere has picked up a fertilization nucleus in the process. a kindly provided by H.-M. Ho;<br />

reprinted from Ho and Chen (1998) with permission of Botanical Bulletin of Academia Sinica.<br />

produced spore of the Oomycota (Chapter 5),<br />

although oospores are also found in the<br />

Monoblepharidales (Chytridiomycota; Fig. 6.25).<br />

In the Oomycota, oospore development begins<br />

with the formation of one or more oospheres<br />

within the larger gametangium, the oogonium.<br />

After fertilization, i.e. the receipt of an antheridial<br />

nucleus by the oosphere, this lays down a<br />

thick wall and becomes the oospore. The number<br />

of oospores per oogonium may vary, and this is<br />

an important taxonomic criterion. Meiotic<br />

nuclear divisions precede oosphere and antheridial<br />

maturation in the Oomycota and nuclear<br />

fusion follows fertilization, so that the oospore is<br />

diploid. The oospore develops a thick outer wall<br />

and lays down food reserves, usually in the form<br />

of lipids. In the Peronosporales the outer wall of<br />

the oospore is surrounded by periplasm, the<br />

residual cy<strong>to</strong>plasm left in the oogonium after<br />

the oospheres have been cleaved out. Oospores<br />

are sedentary (memnospores) and are important<br />

in survival rather than dispersal. They<br />

often require a period of maturation before<br />

germination can occur and may remain dormant<br />

for long periods.<br />

1.4.7 Chlamydospores<br />

In most groups of fungi, terminal or intercalary<br />

segments of the mycelium may become packed<br />

with lipid reserves and develop thick walls<br />

within the original hyphal wall (Fig. 1.22).<br />

The new walls may be colourless or pigmented,<br />

and are often hydrophobic. Structures of<br />

this type have been termed chlamydospores<br />

(Gr. chlamydos ¼ a thick cloak). They are formed<br />

asexually. Generally there is no mechanism for<br />

detachment and dispersal of chlamydospores,<br />

but they may become separated from each other<br />

by the collapse of the hyphae producing them.<br />

They are therefore typical memnospores, forming<br />

important organs of asexual survival, especially<br />

in soil fungi. Chlamydospores may develop<br />

within the sporangiophores of some species<br />

of the Mucorales, e.g. in Mucor racemosus (see<br />

Fig 7.14). The Glomales, which are fungal<br />

partners in symbiotic mycorrhizal associations<br />

with many vascular plants, reproduce primarily<br />

by large, thick-walled chlamydospores. These<br />

develop singly or in clusters (sporocarps) on<br />

coarse hyphae attached <strong>to</strong> their host plants.<br />

They are sedentary in soil but may be dispersed


30 INTRODUCTION<br />

Fig1.22 Chlamydospores formed by soil-borne fungi. (a) Intercalary hyphal chlamydospores in Mucor plumbeus (Zygomycota).<br />

(b) Terminal chlamydospore in Pythium undulatum (Oomycota). Both images <strong>to</strong> same scale.<br />

by wind or by burrowing rodents which eat the<br />

spores. Chlamydospores may also develop within<br />

the multicellular macroconidia of Fusarium spp.<br />

and may survive when other, thin-walled cells<br />

making up the spore are degraded by soil microorganisms.<br />

Similar structures are found in old<br />

hyphae of the aquatic fungus Saprolegnia (see<br />

Fig. 5.6g), either singly or in chains. In this genus,<br />

the chlamydospores may break free from the<br />

mycelium and be dispersed in water currents.<br />

Chlamydospores which are dispersed in this way<br />

are termed gemmae (Lat. gemma ¼ a jewel).<br />

The term chlamydospore is also sometimes<br />

used <strong>to</strong> describe the thick-walled dikaryotic spore<br />

characteristic of smut fungi (Ustilaginales;<br />

Chapter 23) but the term teliospore is preferable<br />

in this context. Hughes (1985) has discussed the<br />

use of the term chlamydospore.<br />

1.4.8 Conidia (conidiospores)<br />

Conidiospores, commonly known as conidia, are<br />

asexual reproductive structures. The word is<br />

derived from the Greek konidion, a diminutive<br />

of konis, meaning dust (Sut<strong>to</strong>n, 1986). Conidia<br />

are found in many different groups of<br />

fungi, but especially within Ascomycota and<br />

Basidiomycota. The term conidium has, unfortunately,<br />

been used in a number of different ways,<br />

so that it no longer has any precise meaning.<br />

It has been defined by Kirk et al. (2001) as<br />

‘a specialized non-motile (cf. zoospore) asexual<br />

spore, usually caducous (i.e. detached), not<br />

developed by cy<strong>to</strong>plasmic cleavage (cf.<br />

sporangiospore) or free cell formation (cf. ascospore);<br />

in certain Oomycota produced through the<br />

incomplete development of zoosporangia which<br />

fall off and germinate <strong>to</strong> produce a germination<br />

tube’. In many fungi conidia represent a means<br />

of rapid spread and colonization from an initial<br />

focus of infection.<br />

In general, conidia are dispersed passively, but<br />

in a few cases discharge is violent. For instance, in<br />

Nigrospora the conidia are discharged by a squirt<br />

mechanism (Webster, 1952), and in Epicoccum<br />

(Fig. 17.8) discharge is brought about by the<br />

rounding-off of a two-ply septum separating the<br />

conidium from its conidiogenous cell (Webster,<br />

1966; Meredith, 1966). In the Helminthosporium<br />

conidial state of Trichometasphaeria turcica, drying<br />

and shrinkage of the conidiophore is associated<br />

with the sudden development of a gas phase,<br />

causing a jolt sufficient <strong>to</strong> project the conidium<br />

(Meredith, 1965; Leach, 1976).<br />

There is great variation in conidial on<strong>to</strong>geny.<br />

This <strong>to</strong>pic will be dealt with more fully<br />

later when considering the conidial states of<br />

Ascomycota, and at this stage it is sufficient <strong>to</strong><br />

distinguish between the major types of conidial<br />

development, which may be either thallic or<br />

blastic. Cells which produce conidia are conidiogenous<br />

cells. The term thallic is used <strong>to</strong> describe<br />

development where there is no enlargement of<br />

the conidium initial (Fig. 1.23a), i.e. the conidium<br />

arises by conversion of a pre-existing segment of<br />

the fungal thallus. An example of this kind is<br />

Galac<strong>to</strong>myces candidus, in which the conidia are


SPORES OF FUNGI<br />

31<br />

Fig1.23 Diagrams <strong>to</strong> illustrate different kinds of conidial development. (a) Thallic development.There is no enlargement of the<br />

conidium initial. (b) Holoblastic development. All the wall layers of the conidiogenous cell balloon out <strong>to</strong> form a conidium initial<br />

recognizably larger than the conidiogenous cell. (c) Enteroblastic tretic development: only the inner wall layers of the conidiogenous<br />

cell are involved in conidium formation.The inner wall layers balloon out through a narrow channel in the outer wall. (d) Phialidic<br />

development: the conidiogenous cell is a phialide.The wall of the phialide is not continuous with the wall surrounding the conidium.<br />

The conidial wall arises de novo from newly synthesized material in the neck of the phialide. Diagrams based on Ellis (1971a).<br />

formed by dissolution of septa along a hypha<br />

(Fig. 10.10). In most conidia, development is<br />

blastic, i.e. there is enlargement of the conidium<br />

initial before it is delimited by a septum.<br />

Two main kinds of blastic development have<br />

been distinguished:<br />

1. Holoblastic, in which both the inner and<br />

outer wall layers of the conidiogenous cell contribute<br />

<strong>to</strong> conidium formation (Fig. 1.23b). An<br />

example of this kind of development is shown<br />

by the conidia of Sclerotinia fructigena (Fig. 15.3).<br />

2. Enteroblastic, in which only the inner wall<br />

layers of the conidiogenous cell are involved in<br />

conidium formation. Where the inner wall layer<br />

balloons out through a narrow pore or channel<br />

in the outer wall layer, development is described<br />

as tretic (Fig. 1.23c). Examples of enteroblastic<br />

tretic development are found in Helminthosporium<br />

velutinum (Fig. 17.12) and Pleospora herbarum<br />

(Fig. 17.9d). Another important method of enteroblastic<br />

development is termed phialidic<br />

development. Here the conidiogenous cell is<br />

a specialized cell termed the phialide. During<br />

the expansion of the first-formed conidium, the<br />

tip of the phialide is ruptured. Further conidia<br />

develop by the extension of cy<strong>to</strong>plasm enclosed<br />

by a new wall layer which is laid down in the<br />

neck of the phialide and is distinct from the<br />

phialide wall. The pro<strong>to</strong>plast of the conidium is<br />

pinched off by the formation of an inwardly<br />

growing flange which closes <strong>to</strong> form a septum<br />

(Fig. 1.23d). New conidia develop beneath the<br />

earlier ones, so that a chain may develop with<br />

the oldest conidium at its apex and the youngest<br />

at its base. Details of phialidic development are<br />

discussed more fully in relation <strong>to</strong> Aspergillus and<br />

Penicillium (p. 299), which reproduce by means of<br />

chains of dry phialoconidia dispersed by wind.<br />

Sticky phialospores which accumulate in slimy<br />

droplets at the tips of the phialides are common<br />

in many genera; they are usually dispersed by<br />

insects, rain splash or other agencies.


32 INTRODUCTION<br />

As mentioned on p. 24, the term conidium is<br />

sometimes used for structures which are probably<br />

homologous <strong>to</strong> sporangia. A series can be<br />

erected in the Peronosporales in which there<br />

are forms with deciduous sporangia which<br />

release zoospores when in contact with water<br />

(e.g. Phy<strong>to</strong>phthora), and other forms which germinate<br />

directly, i.e. by the formation of a germ tube<br />

(e.g. Peronospora). A similar series can be erected<br />

in the Mucorales where in some forms the<br />

number of sporangiospores per sporangium is<br />

reduced <strong>to</strong> several or even one (see Figs. 7.24,<br />

7.26, 7.30). One-spored sporangia may be distinguished<br />

from conidia by being surrounded by<br />

two walls, i.e. that of the sporangium and that<br />

of the spore itself.<br />

There are numerous other kinds of spore<br />

found in fungi, and they are described later in<br />

this book in relation <strong>to</strong> the particular groups in<br />

which they occur.<br />

1.4.9 Anamorphs and teleomorphs<br />

<strong>Fungi</strong> may exist in a range of forms or morphs,<br />

i.e. they may be pleomorphic. The morph which<br />

includes the sexually produced spore form,<br />

e.g. the ascocarp of an ascomycete or the<br />

basidiocarp of a basidiomycete, is termed the<br />

teleomorph (Gr. teleios, teleos ¼ perfect, entire;<br />

morphe ¼ shape, form) (Hennebert & Weresub,<br />

1977). Many fungi also have a morph bearing<br />

asexually produced spores, e.g. conidiomata.<br />

These asexual morphs are termed anamorphs<br />

(Gr. ana ¼ throughout, again, similar <strong>to</strong>). In the<br />

older literature, the term perfect state was used<br />

for the teleomorph and imperfect state for<br />

the anamorph. This is the origin of the name<br />

of the artificial group <strong>Fungi</strong> Imperfecti or<br />

Deuteromycetes, which included fungi believed<br />

<strong>to</strong> reproduce only by asexual means. The term<br />

mi<strong>to</strong>sporic fungi is sometimes used alternatively<br />

for such fungi. The complete range of morphs<br />

belonging <strong>to</strong> any one fungus is termed the<br />

holomorph (Gr. holos ¼ whole, entire) (see<br />

Sugiyama, 1987; Reynolds & Taylor, 1993;<br />

Seifert & Samuels, 2000). Some fungi have more<br />

than one anamorph as in the microconidia and<br />

macroconidia of some Neurospora, Fusarium and<br />

Botrytis species. These distinctive states are<br />

synanamorphs and may play different roles in<br />

the biology of the fungus. The morph may have a<br />

purely sexual role as a fertilizing agent, e.g. in<br />

the case of spermatia of many ascomycetes and<br />

rust fungi. Such states have been termed<br />

andromorphs (Gr. andros ¼ a man, male)<br />

(Parbery, 1996a).<br />

The existence of different states in the life<br />

cycle of a fungus has nomenclatural consequences,<br />

because they had often been described<br />

separately and given different names before the<br />

genetic connection between them was established.<br />

Further, even after the proof of an<br />

anamorph teleomorph relationship, usually<br />

achieved by pure-culture studies, the<br />

anamorphic name may still be in wide use,<br />

especially where it is the more common state<br />

encountered in nature or culture. For example,<br />

most fungal geneticists refer <strong>to</strong> Aspergillus<br />

nidulans (the name of the conidial state) instead<br />

of Emericella nidulans (the name for the ascosporic<br />

state; p. 308). Similarly, most plant pathologists<br />

use Botrytis cinerea, the name for the conidial<br />

state of the fungus causing the common grey<br />

mould disease of many plants, in preference <strong>to</strong><br />

the rarely encountered Sclerotinia (Botryotinia)<br />

fuckeliana, the name given <strong>to</strong> the apothecial<br />

(ascus-bearing) state (see p. 434).<br />

1.5 Taxonomy of fungi<br />

Taxonomy is the science of classification, i.e. the<br />

‘assigning of objects <strong>to</strong> defined categories’ (Kirk<br />

et al., 2001). Classification has three main functions:<br />

it provides a framework of recognizable<br />

features by which an organism under examination<br />

can be identified; it is an attempt <strong>to</strong> group<br />

<strong>to</strong>gether organisms that are related <strong>to</strong> each<br />

other; and it assists in the retrieval of information<br />

about the identified organism in the form of<br />

a list or catalogue.<br />

All taxonomic concepts are man-made and<br />

therefore <strong>to</strong> a certain extent arbitrary. This is<br />

especially true of classical approaches relying<br />

on macroscopic or microscopic observations<br />

because it is a matter of opinion whether the<br />

difference in a particular character say, a spore


TAXONOMY OF FUNGI<br />

33<br />

or the way in which it is formed is significant<br />

<strong>to</strong> distinguish two fungi and, if so, at which<br />

taxonomic level. The great fungal taxonomist<br />

R. W. G. Dennis (1960) described taxonomy as ‘the<br />

art of classifying organisms: not a science but an<br />

art, for its triumphs result not from experiment<br />

but from disciplined imagination guided by<br />

intuition’.<br />

Recently, great efforts have been made at<br />

introducing a seemingly more objective set of<br />

criteria based directly on comparisons of selected<br />

DNA sequences encoding genes with a conserved<br />

biological function, instead of or in addition <strong>to</strong><br />

phenotypic characters. The results of such<br />

comparisons are usually displayed as phylogenetic<br />

trees (see Fig. 1.26), which imply a common<br />

ancestry <strong>to</strong> all organisms situated above a given<br />

branch. Such a grouping is ideally ‘monophyletic’.<br />

However, as we shall see later, quite<br />

different phylogenies may result if different<br />

genes are chosen for comparison. Further, a<br />

decision on the degree of sequence divergence<br />

required for a taxonomic distinction is based<br />

mainly on numerical parameters generated by<br />

elaborate computerized statistical treatments,<br />

occasionally at the expense of sound judgement.<br />

An excessive emphasis on such purely descriptive<br />

studies in the recent literature has led an<br />

eminent mycologist <strong>to</strong> characterize phylogenetic<br />

trees as ‘the most noxious of all weeds’. Despite<br />

their limitations, these methods have led <strong>to</strong> a<br />

revolution in the taxonomy of fungi. At present,<br />

a new, more ‘natural’ classification is beginning<br />

<strong>to</strong> take shape, in which DNA sequence data are<br />

integrated with microscopic, ultrastructural and<br />

biochemical characters. However, many groups<br />

of fungi are still poorly defined, and many more<br />

trees will grow and fall before a comprehensive<br />

taxonomic framework can be expected <strong>to</strong> be in<br />

place. One of the core problems in fungal<br />

taxonomy is the seemingly seamless transition<br />

between the features of two taxa, and the<br />

question as <strong>to</strong> where <strong>to</strong> apply the cut-off point.<br />

To quote Dennis (1960) again, ‘a taxonomic<br />

species cannot exist independently of the<br />

human race; for its constituent individuals can<br />

neither taxonomise themselves in<strong>to</strong> a species,<br />

nor be taxonomised in<strong>to</strong> a species by science in<br />

the abstract; they can only be grouped in<strong>to</strong><br />

species by individual taxonomisers’.<br />

1.5.1 Traditional taxonomic methods<br />

Early philosophers classified matter in<strong>to</strong> three<br />

Kingdoms: Animal, Vegetable, and Mineral.<br />

<strong>Fungi</strong> were placed in the Vegetable Kingdom<br />

because of certain similarities <strong>to</strong> plants such as<br />

their lack of mobility, absorptive nutrition, and<br />

reproduction by spores. Indeed, it was at one<br />

time thought that fungi had evolved from algae<br />

by loss of pho<strong>to</strong>synthetic pigmentation. This was<br />

indicated by the use of such taxonomic groups<br />

as Phycomycetes, literally meaning ‘algal fungi’.<br />

This grouping, approximately synonymous with<br />

the loose term ‘lower fungi’, is no longer used<br />

because it includes taxa not now thought <strong>to</strong><br />

be related <strong>to</strong> each other (chiefly Oomycota,<br />

Chytridiomycota, Zygomycota). Early systems of<br />

classification were based on morphological<br />

(macroscopic) similarity, but the invention of<br />

the light microscope revealed that structures<br />

such as fruit bodies which looked alike could be<br />

ana<strong>to</strong>mically distinct and reproduce in fundamentally<br />

different ways, leading them <strong>to</strong> be<br />

classified apart.<br />

Until the 1980s, the taxonomy of fungi was<br />

based mainly on light microscopic examination<br />

of typical morphological features, giving rise <strong>to</strong><br />

classification schemes which are now known <strong>to</strong><br />

be unnatural. Several examples of unnatural<br />

groups may be found by comparing the present<br />

edition with the previous edition of this textbook<br />

(Webster, 1980). Examples of traditional taxonomic<br />

features include the presence or absence<br />

of septa in hyphae, fine details of the type,<br />

formation and release mechanisms of spores (e.g.<br />

Kendrick, 1971), or aspects of the biology and<br />

ecology of fungi. Useful ultrastructural details,<br />

provided by transmission electron microscopy,<br />

concern the appearance of mi<strong>to</strong>chondria, properties<br />

of the septal pore, details of the cell wall<br />

during spore formation or germination, or the<br />

arrangement of secre<strong>to</strong>ry vesicles in the apex of<br />

growing hyphae (Fig. 1.4). Biochemical methods<br />

have also made valuable contributions, especially<br />

in characterizing higher taxonomic levels.<br />

Examples include the chemical composition of


34 INTRODUCTION<br />

the cell wall (Table 1.1), alternative pathways<br />

of lysine biosynthesis (see p. 67), the occurrence<br />

of pigments (Gill & Steglich, 1987) and the types<br />

and amounts of sugars or polyols (Pfyffer et al.,<br />

1986; Rast & Pfyffer, 1989).<br />

Microscopic features are still important <strong>to</strong>day<br />

for recognizing fungi and making an initial<br />

identification which can then, if necessary, be<br />

backed up by molecular methods. Indeed, the<br />

comparison of DNA sequences obtained from<br />

fungi is meaningful only if these fungi have<br />

previously been characterized and named by<br />

conventional methods. It is therefore just as<br />

necessary <strong>to</strong>day as it ever was <strong>to</strong> teach mycology<br />

students the art of examining and identifying<br />

fungi.<br />

1.5.2 Molecular methods of<br />

fungal taxonomy<br />

A detailed description of modern taxonomic<br />

methods is beyond the scope of this book, and<br />

the reader is referred <strong>to</strong> several in-depth reviews<br />

of the <strong>to</strong>pic (e.g. Kohn, 1992; Clutterbuck, 1995).<br />

A particularly readable introduction <strong>to</strong> this<br />

subject has been written by Berbee and Taylor<br />

(1999). Only the most important molecular<br />

methods are outlined here. They are based<br />

either directly on the DNA sequences or on the<br />

properties of their protein products, especially<br />

enzymes.<br />

Proteins extracted from the cultures of fungi<br />

can be separated by their differential migration<br />

in the electric field of an electrophoresis gel. The<br />

speed of migration is based on the charge and<br />

size of each molecule, resulting in a characteristic<br />

banding pattern. Numerous bands will be<br />

obtained if the electrophoresis gel is stained with<br />

a general protein dye such as Coomassie Blue.<br />

More selective information can be obtained by<br />

isozyme analysis, in which the gel is incubated<br />

in a solution containing a particular substrate<br />

which is converted in<strong>to</strong> a coloured insoluble<br />

product by the appropriate enzyme, or in which<br />

an insoluble substrate such as starch is digested.<br />

In this way, the number and electrophoretic<br />

migration patterns of isoenzymes can be<br />

compared between different fungal isolates.<br />

Protein analysis is useful mainly for<br />

distinguishing different strains of the same<br />

species or members of the same genus (Brasier,<br />

1991a).<br />

Gel electrophoresis can also be used for the<br />

separation of DNA fragments generated by<br />

various methods. One such method is called<br />

RFLP (restriction fragment length polymorphisms)<br />

and involves the digestion of a <strong>to</strong>tal<br />

DNA extract or a previously amplified target<br />

sequence with one or more restriction endonucleases,<br />

i.e. enzymes which cut DNA only at<br />

a particular target site defined by a specific<br />

oligonucleotide sequence. Fragments from this<br />

digest can be blotted from the gel on<strong>to</strong><br />

a membrane; fragments belonging <strong>to</strong> a known<br />

gene can be visualized by hybridizing with<br />

a fluorescent or radioactively labelled DNA<br />

probe of the same gene. In this way, a banding<br />

pattern is obtained and can be compared with<br />

that of other fungal isolates prepared under<br />

identical experimental conditions.<br />

A similar method, RAPD (random amplified<br />

polymorphic DNA), produces DNA bands not by<br />

digestion, but by the amplification of DNA<br />

sequences. For this purpose, a DNA extract is<br />

incubated with a DNA polymerase, deoxynucleoside<br />

triphosphates and one or more short<br />

oligonucleotides which act as primers for the<br />

polymerase by binding <strong>to</strong> complementary DNA<br />

sequences which should be scattered throughout<br />

the genome. Amplification is achieved by means<br />

of the PCR (polymerase chain reaction), in which<br />

the mixture is subjected <strong>to</strong> repeated cycles of<br />

different temperatures suitable for annealing of<br />

DNA and primer, polymerization, and dissociation<br />

of double-stranded DNA. The largest possible<br />

size of the amplification product depends on the<br />

polymerization time; bands visible on a gel will<br />

be produced only if two primer binding sites<br />

happen <strong>to</strong> be in close proximity <strong>to</strong> each other, so<br />

that the intervening stretch of DNA sequence can<br />

be amplified from both ends within the chosen<br />

polymerization time. The number and size of<br />

RAPD bands on electrophoresis gels can be<br />

compared between different fungi, provided<br />

that all samples have been produced under<br />

identical conditions.<br />

Isozyme, RFLP and RAPD analyses all generate<br />

data which are useful mainly for comparing


TAXONOMY OF FUNGI<br />

35<br />

closely related isolates. Since the results strongly<br />

depend on the experimental conditions<br />

employed, there are no universal databases for<br />

these types of analysis. Further, they are unsuitable<br />

for comparisons of distantly related or<br />

unrelated organisms. A breakthrough in the<br />

taxonomy of fungi as well as other organisms<br />

was achieved when primers were developed<br />

which guided the PCR amplification of specific<br />

stretches of DNA universally present and fulfilling<br />

a homologous function in all life forms. Once<br />

amplified, the sequence of bases can be determined<br />

easily. Such methods were first applied <strong>to</strong><br />

bacterial systematics with spectacular results<br />

(Woese, 1987). In eukaryotes, the most widely<br />

used target sequences are those encoding the 18S<br />

or 28S ribosomal RNA (rRNA) molecules, which<br />

fulfil a structural role in the small or large<br />

ribosomal subunits (respectively), or the noncoding<br />

DNA stretches (ITS, internal transcribed<br />

spacers), which physically separate these genes<br />

from each other and from the 5.8S rRNA<br />

sequence in the nuclear genome (Fig. 1.24;<br />

White et al., 1990). The structural role which<br />

rRNA molecules play in the assembly of ribosomes<br />

requires them <strong>to</strong> take up a particular<br />

configuration which is stable because of intramolecular<br />

base-pairing. Since certain regions of<br />

each rRNA molecule hybridize with complementary<br />

regions within the same molecule or with<br />

other rRNA molecules, mutations in the DNA<br />

encoding these regions are rare because they<br />

would impair hybridization and thus the functioning<br />

of the rRNA molecule unless accompanied<br />

by a mutation at the complementary<br />

binding site. The non-pairing loop regions of<br />

the rRNA gene and the ITS sequences are not<br />

subjected <strong>to</strong> such a strong selective pressure and<br />

thus tend <strong>to</strong> show a higher rate of mutation.<br />

Nucleotide sequences therefore permit the<br />

comparison of closely related species or even<br />

strains of the same species (ITS sequences), as<br />

well as that of distantly related taxa or even<br />

members of different kingdoms (18S or 28S<br />

rRNA). Further, because extensive databases are<br />

now available, the sequence analysis of a single<br />

fungus can provide meaningful taxonomic information<br />

when compared with existing sequences.<br />

In addition <strong>to</strong> ribosomal DNA sequences, genes<br />

encoding cy<strong>to</strong>chrome oxidase (COX), tubulins or<br />

other proteins with conserved functions are now<br />

used extensively for phylogenetic purposes.<br />

Once comparative data have been obtained<br />

either by banding patterns or gene sequencing,<br />

they need <strong>to</strong> be evaluated. This is usually done by<br />

converting the data in<strong>to</strong> a matrix, e.g. by scoring<br />

the absence or presence of a particular band.<br />

With comparisons of aligned DNA sequences,<br />

only informative positions are selected for the<br />

matrix, i.e. where variations in the nucleotides<br />

between different fungi under investigation are<br />

observed. When the matrix has been completed,<br />

it can be subjected <strong>to</strong> statistical treatments,<br />

and phylogenetic trees are drawn by a range of<br />

algorithms. In some, the degree of relatedness<br />

of taxa is indicated by the length of the branch<br />

separating them (see Figs. 1.25, 1.26). Such<br />

information is thought <strong>to</strong> be of evolutionary<br />

significance; the greater the number of differences<br />

between two organisms, the earlier the<br />

separation of their evolutionary lines should<br />

have occurred.<br />

1.5.3 How old are fungi?<br />

Several lines of evidence indicate that fungi are a<br />

very ancient group of organisms. Berbee and<br />

Taylor (2001) have attempted <strong>to</strong> add a timescale<br />

<strong>to</strong> phylogenetic trees by applying the concept of<br />

a ‘molecular clock’, i.e. the assumption that the<br />

rate of mutations leading <strong>to</strong> phylogenetic diversity<br />

is constant over time and in various groups<br />

of organisms. By calibrating their molecular<br />

clock against fossil evidence, Berbee and Taylor<br />

(2001) estimated that fungi may have separated<br />

from animals some 900 million years ago,<br />

i.e. long before the evolution of terrestrial<br />

organisms. This estimate is consistent with the<br />

discovery of fossilized anas<strong>to</strong>mozing hypha-like<br />

structures in sediments about 1 billion years<br />

old (Butterfield, 2005). <strong>Fungi</strong> recognizable as<br />

Chytridiomycota, Zygomycota and Ascomycota<br />

have been discovered among fossils of early<br />

terrestrial plants from the Lower Devonian<br />

Rhynie chert, formed some 400 million years<br />

ago (Taylor et al., 1992, 1999, 2005). It is apparent<br />

that these early terrestrial plants already entertained<br />

mycorrhizal symbiotic associations


36 INTRODUCTION<br />

Fig1.24 The spatial arrangement of a nuclear rRNA gene repeat unit. Each haploid fungal genome contains about 50 250 copies<br />

of this repeat, depending on the species (Vilgalys & Gonzalez,1990).The three structural rRNA genes encoded by one repeat unit,<br />

i.e.18S, 5.8S and 28S, are separated by internal and external transcribed spacers (ITS and ETS, respectively). Adjacent copies of<br />

the repeat unit are separated by a short non-transcribed spacer (NTS).The whole unit is transcribed in<strong>to</strong> a 45S precursor RNA in<br />

one piece, followed by excision of the three structural RNA molecules from the spacers which are not used.The 5S rRNA gene<br />

is encoded at a separate locus.The18S rRNA molecule is part of the small ribosomal subunit, whereas the other three contribute<br />

<strong>to</strong> the large subunit.<br />

Fig1.25 The phylogenetic relationships of <strong>Fungi</strong> and fungus-like organisms studied by mycologists (printed in bold), with other<br />

groups of Eukaryota.The analysis is based on comparisons of18SrDNA sequences.Modified andredrawn from Brunset al.(1991)and<br />

Berbee and Taylor (1999).


TAXONOMY OF FUNGI<br />

37<br />

Fig1.26 Phylogenetic relationships within the Eumycota, based on18S rDNA comparisons.This tree illustrates the analytical<br />

power of molecular phylogenetic analyses; all four phyla of Eumycota are resolved. However, it also highlights problems in that<br />

Basidiobolus groups with the Chytridiomycota, although sharing essential biological features with the Zygomycota, and that<br />

conversely Blas<strong>to</strong>cladiella groups with the Zygomycota instead of the Chytridiomycota. Modified and redrawn from Berbee and<br />

Taylor (2001), with kind permission of Springer Science and Business media.<br />

with glomalean members of the Zygomycota<br />

(see p. 218).<br />

1.5.4 The taxonomic system adopted<br />

in this book<br />

The discipline of fungal taxonomy is evolving at<br />

an unprecedented speed at present due mainly<br />

<strong>to</strong> the contributions of molecular phylogeny.<br />

Numerous taxonomic systems exist, but this is<br />

not the place <strong>to</strong> discuss their relative merits<br />

(see Whittaker, 1969; Margulis et al., 1990;<br />

Alexopoulos et al., 1996; Cavalier-Smith, 2001;<br />

Kirk et al., 2001). In this book we have tried <strong>to</strong><br />

follow the classification proposed in The Mycota<br />

Volumes VIIA and VIIB (McLaughlin et al., 2001),<br />

but even in these volumes the authors of<br />

different chapters have used their own favoured<br />

systems of classification rather than adopting<br />

an imposed one. In cases of doubt, we have<br />

attempted <strong>to</strong> let clarity prevail over pedantry.<br />

<strong>Fungi</strong> in the widest sense, as organisms<br />

traditionally studied by mycologists, currently<br />

fall in<strong>to</strong> three kingdoms of Eukaryota, i.e. the<br />

Eumycota which contain only fungi, and the<br />

Pro<strong>to</strong>zoa and Chromista (¼ Straminipila), both<br />

of which contain mainly organisms not studied<br />

by mycologists and were formerly lumped<br />

<strong>to</strong>gether under the name Pro<strong>to</strong>ctista (Beakes,<br />

1998; Kirk et al., 2001). The Pro<strong>to</strong>zoa are<br />

no<strong>to</strong>riously difficult <strong>to</strong> resolve by phylogenetic<br />

means, and the only firm statement which can<br />

be made at present is that they are a diverse<br />

and ancient group somewhere between the<br />

higher Eukaryota (‘crown eukaryotes’) and the


38 INTRODUCTION<br />

Table1.2. The classification scheme adopted in<br />

this book, showing mainly those groups treated in<br />

some detail.<br />

KINGDOM PROTOZOA<br />

Myxomycota (Chapter 2)<br />

Acrasiomycetes<br />

Dictyosteliomycetes<br />

Pro<strong>to</strong>steliomycetes<br />

Myxomycetes<br />

Plasmodiophoromycota (Chapter 3)<br />

Plasmodiophorales<br />

Hap<strong>to</strong>glossales (Oomycota?)<br />

KINGDOM STRAMINIPILA<br />

Hyphochytriomycota (Chapter 4)<br />

Labyrinthulomycota (Chapter 4)<br />

Labyrinthulomycetes<br />

Thraus<strong>to</strong>chytriomycetes<br />

Oomycota (Chapter 5)<br />

Saprolegniales<br />

Pythiales<br />

Peronosporales<br />

KINGDOM FUNGI (EUMYCOTA)<br />

Chytridiomycota (Chapter 6)<br />

Chytridiomycetes<br />

Zygomycota (Chapter 7)<br />

Zygomycetes<br />

Trichomycetes<br />

Ascomycota (Chapter 8)<br />

Archiascomycetes (Chapter 9)<br />

Hemiascomycetes (Chapter10)<br />

Plec<strong>to</strong>mycetes (Chapter11)<br />

Hymenoascomycetes<br />

Pyrenomycetes (Chapter12)<br />

Erysiphales (Chapter13)<br />

Pezizales (Chapter14)<br />

Helotiales (Chapter15)<br />

Lecanorales/lichens (Chapter16)<br />

Loculoascomycetes (Chapter17)<br />

Basidiomycota (Chapter18)<br />

Homobasidiomycetes (Chapter19)<br />

Homobasidiomycetes: gasteromycetes<br />

(Chapter 20)<br />

Heterobasidiomycetes (Chapter 21)<br />

Urediniomycetes (Chapter 22)<br />

Ustilaginomycetes (Chapter 23)<br />

prokaryotes (Kumar & Rzhetsky, 1996). An overview<br />

of eukaryotic organisms, in which those<br />

groups treated in this book are highlighted,<br />

is given in Fig. 1.25. Among the Pro<strong>to</strong>zoa, the<br />

Plasmodiophoromycota are given extensive<br />

treatment because of their role as pathogens<br />

of plants (Chapter 3), whereas the various<br />

forms of slime moulds are considered only<br />

briefly (Chapter 2). Similarly brief overviews<br />

will be given of most groups of Straminipila<br />

studied by mycologists (Chapter 4), except for<br />

the Oomycota which, despite their separate<br />

evolutionary origin, represent a major area of<br />

mycology (Chapter 5). All remaining chapters<br />

deal with members of the Eumycota (¼ Kingdom<br />

<strong>Fungi</strong>). The scheme is summarized in Table 1.2<br />

and illustrated in Fig. 1.26. An overview of the<br />

nomenclature used for describing taxa within<br />

the Eumycota is given in Table 1.3.<br />

In the past, fungi which solely or mainly<br />

reproduce asexually (<strong>Fungi</strong> Imperfecti,<br />

Deuteromycota, mi<strong>to</strong>sporic fungi, anamorphic<br />

fungi) were considered separately from their<br />

sexually reproducing relatives the teleomorphs,<br />

and separate anamorph and teleomorph<br />

genera were erected. However, information<br />

from pure-culture studies and molecular phylogenetic<br />

approaches has linked many anamorphs<br />

with their teleomorphs. For instance, the conidial<br />

(imperfect) state of the common brownrot<br />

fungus of apples and other fruits is called<br />

Monilia fructigena, whereas the sexual (perfect)<br />

Table 1.3. Example of the hierarchy of taxonomic<br />

terms. The wheat stem rust fungus, Puccinia<br />

graminis, is used as an example.<br />

Kingdom <strong>Fungi</strong><br />

Subkingdom Eumycota<br />

Phylum Basidiomycota<br />

Class Urediniomycetes<br />

Order Uredinales<br />

Family Pucciniaceae<br />

Genus Puccinia<br />

Species Puccinia graminis<br />

Race Puccinia graminis<br />

f. sp. tritici


TAXONOMY OF FUNGI<br />

39<br />

state is apothecial, being called Sclerotinia<br />

(Monilinia) fructigena. As far as is possible, we<br />

shall consider anamorphic states of fungi in<br />

the context of their known sexual state. Thus,<br />

an account of the brown-rot of fruits, although<br />

encountered predominantly as the conidial<br />

state, will be given in the chapter dealing<br />

with apothecial fungi (Helotiales, Chapter 15).<br />

Where practical, we have given the teleomorph<br />

name priority over the anamorph. As a longterm<br />

future goal, Seifert and Samuels (2000)<br />

and Seifert and Gams (2001) have outlined<br />

a unified taxonomy which might ultimately<br />

lead <strong>to</strong> the abolition of the names of anamorphic<br />

genera.<br />

However, with certain ecological groups<br />

such as the Ingoldian aquatic fungi (Section<br />

25.2) and nema<strong>to</strong>phagous fungi (Section 25.1),<br />

which have diverse relationships, we have<br />

deliberately chosen <strong>to</strong> consider them in their<br />

ecological context rather than along with their<br />

varied taxonomic relatives.


2<br />

Pro<strong>to</strong>zoa: Myxomycota (slime moulds)<br />

2.1 <strong>Introduction</strong><br />

When the first slime moulds were described<br />

by Johann H. F. Link in 1833, they were given<br />

the term myxomycetes (Gr. myxa ¼ slime). Link<br />

used the suffix -mycetes because of the superficial<br />

similarity of the fructifications of slime moulds<br />

with the fruit bodies of certain fungi, notably<br />

Gasteromycetes (see Chapter 20). Although it<br />

has been appreciated for some time that they<br />

lack any true relationship with the Eumycota<br />

(de Bary, 1887; Whittaker, 1969), slime moulds<br />

have none the less been studied mainly by<br />

mycologists rather than pro<strong>to</strong>zoologists, probably<br />

because they occur in the same habitats<br />

as fungi and are routinely encountered during<br />

fungus forays. Since slime moulds are only<br />

rarely covered by zoology courses even <strong>to</strong>day,<br />

they are briefly described in this chapter,<br />

referring <strong>to</strong> more specialized literature as<br />

appropriate.<br />

Slime moulds differ substantially from<br />

the Eumycota not only in phylogenetic terms,<br />

but also regarding their physiology and ecology.<br />

Their vegetative state is that of individual<br />

amoebae in the cellular slime moulds, or of<br />

a multinuclear (coenocytic) plasmodium in the<br />

plasmodial slime moulds. Motile stages bearing<br />

usually two anterior whiplash-type flagella may<br />

be present in the plasmodial slime moulds<br />

(Sections 2.4, 2.5) and in the Plasmodiophoromycota<br />

(Chapter 3). Amoebae or plasmodia feed<br />

by the ingestion (phagocy<strong>to</strong>sis) of bacteria,<br />

yeast cells or other amoebae. This is followed by<br />

intracellular digestion in vacuoles. The mode of<br />

nutrition in slime moulds is therefore fundamentally<br />

different from extracellular degradation<br />

and absorption as shown by Eumycota.<br />

Numerous phylogenetic analyses of DNA<br />

sequences encoding rRNA molecules and various<br />

structural proteins or enzymes have been carried<br />

out, but the results obtained are difficult <strong>to</strong><br />

interpret because the comparison of different<br />

genes have led <strong>to</strong> rather variable phylogenetic<br />

schemes. Of the four groups treated in this chapter,<br />

it seems that the Dictyosteliomycetes,<br />

Pro<strong>to</strong>steliomycetes and Myxomycetes are related<br />

<strong>to</strong> each other whereas the Acrasiomycetes have<br />

a different evolutionary origin (Baldauf, 1999;<br />

Baldauf et al., 2000). The general evolutionary<br />

background is, however, still rather diffuse in<br />

these lower eukaryotes.<br />

2.2 Acrasiomycetes: acrasid cellular<br />

slime moulds<br />

The Acrasiomycetes, or Acrasea as they are called<br />

in zoological classification schemes, are a small<br />

group currently comprising 12 species in six<br />

genera (Kirk et al., 2001). Although appearing<br />

somewhat removed from the bulk of the slime<br />

moulds, they still clearly belong <strong>to</strong> the Pro<strong>to</strong>zoa<br />

(Roger et al., 1996). The trophic stage consists<br />

of amoebae which are morphologically distinct<br />

from those of the dictyostelid cellular slime<br />

moulds (Section 2.3) in having a cylindrical,<br />

rather than flattened, body bearing a single


DICTYOSTELIOMYCETES: DICTYOSTELID SLIME MOULDS<br />

41<br />

Fig 2.1 Amoebae of cellular slime moulds.The arrows indicate the direction of movement at the time when the pho<strong>to</strong>micrographs<br />

were taken. (a) Limax-type amoeba of Acrasisrosea, an acrasid cellular slime mould.Note the absence of granular contents from the<br />

lobose pseudopodium at the tip of the amoeba. (b) Amoeba of Pro<strong>to</strong>stelium mycophaga with filose pseudopodia. Reproduced from<br />

Zuppinger and Roos (1997), with permission from Elsevier; original prints kindly supplied by C. Zuppinger.<br />

large-lobed (lobose) anterior pseudopodium. The<br />

granular cellular contents trail behind the pseudopodium,<br />

which appears clear. The posterior<br />

end is knob-shaped and is called the uroid<br />

(Fig. 2.1a). Such amoebae are of the limax type<br />

because their movement resembles that of slugs<br />

of the genus Limax. Good accounts of the acrasids<br />

have been given by Olive (1975) and Blan<strong>to</strong>n<br />

(1990).<br />

Acrasid slime moulds are common on decaying<br />

plant matter, in soil, on dung and on rotting<br />

mushrooms, but they are rarely recorded<br />

because of their small size, which necessitates<br />

observations with a dissecting microscope. The<br />

most readily recognized species is Acrasis rosea,<br />

which has orange- or pink-coloured amoebae<br />

due <strong>to</strong> the presence of carotenoid pigments,<br />

including <strong>to</strong>rulene (Fuller & Rakatansky, 1966).<br />

Acrasis rosea can be observed if dead twigs,<br />

leaves or fruits are incubated on weak nutrient<br />

agar for a few days. Spore-bearing structures<br />

called sorocarps (Gr. sorus ¼ heap, karpos ¼ fruit)<br />

will develop, and spores can be transferred <strong>to</strong><br />

fresh agar with yeast cells as a food source<br />

(Blan<strong>to</strong>n, 1990). The uninucleate amoebae feed<br />

on yeast cells, bacteria or fungal spores and can<br />

encyst under unfavourable conditions, especially<br />

drought, <strong>to</strong> form microcysts. Each microcycst<br />

germinates again <strong>to</strong> release a single amoeba.<br />

Eventually amoebae aggregate <strong>to</strong> form a pseudoplasmodium,<br />

in which the individual amoebae<br />

retain their identity but are surrounded by a<br />

common sheath. The chemical signal for aggregation<br />

is unknown but it is not cyclic AMP (cAMP)<br />

as in the dictyostelid slime moulds (see below).<br />

The pseudoplasmodium develops in<strong>to</strong> a<br />

branched sorocarp in which the amoebae align<br />

themselves in single rows and then round off,<br />

each forming a walled spore. Each spore germinates<br />

<strong>to</strong> release a single amoeba. The cells<br />

making up the stalk of the sorocarp also encyst<br />

and are capable of germination (Olive, 1975).<br />

Sexual reproduction in the acrasid slime<br />

moulds is unknown.<br />

2.3 Dictyosteliomycetes:<br />

dictyostelid slime moulds<br />

The Dictyosteliomycetes (zool.: Dictyostelia) are<br />

a group of cellular slime moulds comprising


42 PROTOZOA: MYXOMYCOTA (SLIME MOULDS)<br />

46 species in four genera (Kirk et al., 2001).<br />

The best-known example is Dictyostelium which<br />

has been so named because the stalk of its multicellular<br />

sorocarp appears as a network, made up<br />

from cellulose walls secreted by the amoebae<br />

from which it is formed. Dictyostelium spp. are<br />

common in soil, on decaying plant material<br />

and on dung, and can be demonstrated by smearing<br />

non-nutrient agar with cells of a suitable<br />

bacterial food such as Escherichia coli or Klebsiella<br />

aerogenes, and adding a small crumb of moistened<br />

soil <strong>to</strong> the centre of the bacterial smear.<br />

Amoebae will creep out of the soil and consume<br />

the bacteria. At the end of the feeding<br />

phase, sorocarps develop and isolations can be<br />

made (Cavender, 1990). An axenic defined<br />

medium has been developed for D. discoideum<br />

and has greatly facilitated experimentation<br />

with this organism (Franke & Kessin, 1977).<br />

Good general accounts of the dictyostelids are<br />

those by K. B. Raper (1984), Cavender (1990) and<br />

Alexopoulos et al. (1996). The his<strong>to</strong>ry of research<br />

on Dictyostelium has been recounted by Bonner<br />

(1999). Work on D. discoideum has contributed<br />

significantly <strong>to</strong> our understanding of the key<br />

features of eukaryotic cell biology, especially<br />

signalling events, phagocy<strong>to</strong>sis, and the evolution<br />

of multicellularity in animals. Consequently,<br />

there is a vast literature on this organism.<br />

An excellent introduction <strong>to</strong> the impact of<br />

research on D. discoideum on general eukaryotic<br />

biology is the book by Kessin (2001), and challenging<br />

questions have been summarized by Ratner<br />

and Kessin (2000). Bonner (2001) has also<br />

provided a stimulating read.<br />

The life cycle of D. discoideum is shown in<br />

Fig. 2.2. Amoebae of dictyostelids are morphologically<br />

different from those of acrasids in that<br />

they have filose (acutely pointed) rather than<br />

lobose pseudopodia (see Fig. 2.1b). Each spore<br />

from a sorocarp germinates <strong>to</strong> give rise <strong>to</strong> one<br />

uninucleate haploid amoeba which feeds by<br />

phagocy<strong>to</strong>sis of bacteria. Amoebae reproduce<br />

asexually by division <strong>to</strong> form two haploid daughter<br />

amoebae. As with acrasid slime moulds,<br />

the amoebae of dictyostelids can form microcysts<br />

under unfavourable environmental conditions.<br />

Encystment may be triggered by the production<br />

of ammonia, which thus functions as a<br />

signal molecule (Cotter et al., 1992). Sexual reproduction<br />

occurs by means of macrocysts and is<br />

initiated when two compatible amoebae meet<br />

and fuse. Both homothallic and heterothallic<br />

species and strains of Dicy<strong>to</strong>stelium are known.<br />

In D. discoideum, fusion is inhibited by light and<br />

by the presence of cAMP, but is stimulated by<br />

ethylene (Amagai, 1992). The fusion cell is greatly<br />

enlarged relative <strong>to</strong> the two progeni<strong>to</strong>r amoebae.<br />

This giant cell attracts unfused amoebae which<br />

aggregate and secrete a sheath (primary wall)<br />

around themselves and the zygote. Inside the<br />

primary wall, the giant cell undergoes karyogamy,<br />

and the resulting zygote feeds cannibalistically<br />

on the other amoebae by phagocy<strong>to</strong>sis<br />

and eventually produces a secondary wall.<br />

Cellulose seems <strong>to</strong> be the main structural wall<br />

polymer. Meiosis is followed by mi<strong>to</strong>tic divisions<br />

and cy<strong>to</strong>plasmic cleavage, and the macrocyst<br />

germinates <strong>to</strong> release numerous haploid uninucleate<br />

amoebae (Nickerson & Raper, 1973;<br />

Szabo et al., 1982).<br />

The most striking feature of D. discoideum<br />

is the aggregation of thousands of amoebae <strong>to</strong><br />

form a pseudoplasmodium with radiating arms<br />

(Figs. 2.3a,b). This is a vegetative process not<br />

involving meiosis or mi<strong>to</strong>sis. Aggregation is<br />

initiated when the bacterial food supply is<br />

exhausted, and follows the gradient of a hormone<br />

which causes directional (chemotactic)<br />

movement of starving amoebae (Konijn et al.,<br />

1967; Swanson & Taylor, 1982). In the case of<br />

D. discoideum, the hormone is cAMP (Konijn<br />

et al., 1967), but other molecules are implicated<br />

in this role in different dictyostelids. Upon exposure<br />

<strong>to</strong> a cAMP gradient, amoebae of D. discoideum<br />

change their shape from isodiametric <strong>to</strong> elongated,<br />

with the migrating tip pointing <strong>to</strong>wards<br />

the highest cAMP concentration. Migration<br />

occurs in waves which correspond <strong>to</strong> the production<br />

of cAMP by starving amoebae, its detection<br />

and further synthesis by neighbouring amoebae,<br />

and its degradation by cAMP phosphodiesterase<br />

(Nagano, 2000; Weijer, 2004). In this way, waves<br />

of cAMP diffuse outwards, and waves of amoebae<br />

migrate inwards. During aggregation, amoebae<br />

migrate <strong>to</strong> the centre or one of the arms of the<br />

pseudoplasmodium. This is a highly co-ordinated<br />

effort in which hundreds of thousands of


DICTYOSTELIOMYCETES: DICTYOSTELID SLIME MOULDS<br />

43<br />

Fig 2.2 Life cycle of Dictyostelium discoideum.The central feature is the haploid amoeba which is free-living in the soil. It divides<br />

mi<strong>to</strong>tically <strong>to</strong> produce two daughter amoebae or, under unfavourable conditions, may form a microcyst. If two amoebae of<br />

compatible mating type meet, a diploid macrocyst may be formed which can remain dormant for some time and eventually<br />

germinates by meiosis and then mi<strong>to</strong>sis <strong>to</strong> release numerous haploid amoebae.Under certain circumstances, starvation may lead <strong>to</strong><br />

aggregation of amoebae <strong>to</strong> form a slug and a sorocarp in which individual amoebae become converted in<strong>to</strong> spores.These are purely<br />

asexual, and meiosis is not involved in their formation or germination.Open and closed circles represent haploid nuclei of opposite<br />

mating type; diploid nuclei are larger and half-filled. Key events in the life cycle are plasmogamy (P), karyogamy (K) and meiosis (M).<br />

amoebae from an area of 1 cm 2 of soil can be<br />

involved. Aggregating amoebae adhere <strong>to</strong> each<br />

other and secrete a common slime sheath<br />

(Figs. 2.3c,d). Eventually they pile up <strong>to</strong> form a<br />

compact bullet-shaped slug which flops over<br />

on<strong>to</strong> the substratum. In D. discoideum and some<br />

other species, the slug undergoes a period of<br />

migration <strong>to</strong>wards the light (Figs. 2.3e g). The<br />

individuality of amoebae is retained within the<br />

slug. As the slug moves along, it leaves behind a<br />

slime trail. Within the slug, the amoebae are<br />

divided in<strong>to</strong> two functionally different populations,<br />

i.e. an anterior group of large, highly<br />

vacuolated cells (pre-stalk cells) and a posterior<br />

group of smaller ones, the pre-spore cells<br />

(Fig. 2.4). It is the pre-stalk group of cells which<br />

co-ordinates slug migration by secreting cAMP.<br />

Various environmental stimuli can direct movement.<br />

For instance, the anterior end of the slug<br />

follows an oxygen gradient but is repelled by<br />

ammonia. Temperature as well as light can also<br />

act as triggers of directed movement. The end of<br />

the migration phase is marked by the roundingoff<br />

and erection of the pseudoplasmodium <strong>to</strong><br />

form a flat-based, somewhat conical structure,<br />

which undergoes further development by differentiating<br />

in<strong>to</strong> a multicellular stalk composed of<br />

the large anterior cells, and the sorus which rises<br />

up on the outside of the stalk (Figs. 2.3h j, 2.4).<br />

This final stage of development is called culmination.<br />

About 80% of the amoebae become<br />

converted in<strong>to</strong> spores, with the remainder<br />

being sacrificed for the formation of the fruit<br />

body structure.


44 PROTOZOA: MYXOMYCOTA (SLIME MOULDS)<br />

Fig 2.3 Dictyostelium discoideum development. (a) Aggregation of amoebae. (b) Aggregation, enlarged. (c) Amoebae feeding<br />

on bacteria; note their isodiametric shape. (d) Aggregating amoebae; note their elongated shape. (e) Late aggregation stage.<br />

(f,g) Migration stage. (h) Culmination; the spore mass is rising around the stalk. (i) Spore mass almost at the apex of the stalk.<br />

(j) Mature sorocarps.<br />

The ability of free-living individual amoebae<br />

of Dictyostelium <strong>to</strong> aggregate in<strong>to</strong> the multicellular<br />

slug has led <strong>to</strong> dictyostelid slime moulds<br />

being called social amoebae (Kessin, 2001). This<br />

phenomenon gives rise <strong>to</strong> interesting and fundamental<br />

questions. To give an example, since<br />

amoebae in the anterior end of the slug become<br />

stalk cells and are thus excluded from perpetuation<br />

as spores, cells skiving off <strong>to</strong> the rear of the<br />

slug and thereby avoiding self-sacrifice would<br />

have a selective advantage. ‘Cheater strains’ are<br />

indeed known from nature and the labora<strong>to</strong>ry;


PROTOSTELIOMYCETES: PROTOSTELID PLASMODIAL SLIME MOULDS<br />

45<br />

Fig 2.4 Dictyostelium discoideum. Development of sorocarp (after Bonner,1944). (a) (c) Aggregation. (d) (h) Migration.<br />

(i) (n) Culmination. C 1<br />

End of aggregation. H 1<br />

End of migration. I 1<br />

Beginning of culmination and stalk formation. J 1<br />

Flattened stage<br />

of culmination. I 1<br />

A later stage of culmination.<br />

some of them cheat only <strong>to</strong> a degree or only<br />

if altruistic non-cheater strains are present,<br />

whereas others are entirely unable <strong>to</strong> make a<br />

fruit body in the absence of wild-type amoebae<br />

prepared <strong>to</strong> form the pre-stalk cells (Dao et al.,<br />

2000; Strassmann et al., 2000). The cheater<br />

phenomenon has raised thought-provoking questions<br />

about the evolution and control of cheating<br />

in social systems (Hudson et al., 2002).<br />

Another interesting aspect involves the mode<br />

of nutrition of Dictyostelium by the phagocy<strong>to</strong>sis<br />

of bacterial cells. Several bacteria pathogenic <strong>to</strong><br />

humans and other animals, e.g. Pseudomonas<br />

aeruginosa and Legionella pneumophila, also kill<br />

Dictyostelium upon ingestion (Solomon et al.,<br />

2000; Pukatzki et al., 2002). The observation<br />

that interactions between Dictyostelium amoebae<br />

and phagocy<strong>to</strong>sed bacterial pathogens are similar<br />

<strong>to</strong> those involving human phagocytes may<br />

stimulate further research on this fascinating<br />

slime mould (Steinert & Heuner, 2005).<br />

2.4 Pro<strong>to</strong>steliomycetes: pro<strong>to</strong>stelid<br />

plasmodial slime moulds<br />

This class of organisms (zool.: Pro<strong>to</strong>stelea) comprises<br />

14 genera and 35 species (Kirk et al., 2001).


46 PROTOZOA: MYXOMYCOTA (SLIME MOULDS)<br />

Useful treatments of the group have been<br />

written by Olive (1967, 1975) and Spiegel (1990).<br />

Pro<strong>to</strong>stelids are ubiqui<strong>to</strong>us on decaying plant<br />

parts in soil and humus, as well as on dung<br />

or in freshwater. They occur in all climatic<br />

zones from the tundra <strong>to</strong> tropical rainforests.<br />

Pro<strong>to</strong>stelids produce amoebae with filose pseudopodia<br />

(Fig. 2.1b), feeding phagocy<strong>to</strong>tically on<br />

bacteria, yeast cells or spores of fungi. Some<br />

species also produce small plasmodia, thereby<br />

providing structural affinities <strong>to</strong> both the cellular<br />

and plasmodial slime moulds. Sporulation<br />

occurs by the conversion of a feeding amoeba<br />

or plasmodium in<strong>to</strong> a round prespore cell<br />

which then rises at the tip of a delicate acellular<br />

stalk, ultimately forming one or several spores<br />

in a single sporangium. It is possible <strong>to</strong> isolate<br />

pro<strong>to</strong>stelids by transferring a spore from its<br />

stalk on<strong>to</strong> a weak nutrient agar plate with<br />

appropriate food organisms.<br />

Pro<strong>to</strong>stelium is a typical member of the group<br />

(Fig. 2.5). The sporocarp consists of a long, slender<br />

stalk about 75 mm long, bearing a single spherical<br />

spore about 4 10 mm in diameter. The spore<br />

is deciduous and readily detached. Upon germination,<br />

a single uninucleate amoeba with thin<br />

pseudopodia emerges. The amoeboid stage<br />

feeds voraciously on yeast cells and may also<br />

feed cannibalistically on amoebae of the same<br />

species. Development of the sporocarp probably<br />

follows the generalized pattern described by<br />

Olive (1967) and summarized in Fig. 2.6. When<br />

feeding s<strong>to</strong>ps, the amoeba rounds off and<br />

heaps its pro<strong>to</strong>plasm in the centre <strong>to</strong> form the<br />

‘hat-shaped’ stage (Fig. 2.6b). A membranous,<br />

pliable, impermeable sheath develops over the<br />

surface of the cell. When the pro<strong>to</strong>plast contracts<br />

in<strong>to</strong> the central hump, the sheath collapses<br />

at the margins, forming the disc-like base <strong>to</strong><br />

the stalk of the sporocarp. This may be the<br />

structural equivalent of the hypothallus of the<br />

Myxomycetes (see p. 48). Within the pro<strong>to</strong>plast,<br />

a granular basal core, the steliogen, differentiates<br />

and begins <strong>to</strong> mould a hollow tube<br />

(Figs. 2.6d,e). As the tube extends at its tip,<br />

the pro<strong>to</strong>plast migrates upwards, always seated<br />

on <strong>to</strong>p of the growing tip. The entire structure<br />

remains covered by the sheath. Tube extension<br />

is an actin myosin-driven process (Spiegel<br />

et al., 1979). Ultimately, the steliogen is left<br />

behind at the tip of the stalk <strong>to</strong> form an apophysis<br />

(Fig. 2.5a), and the pro<strong>to</strong>plast secretes a cell<br />

wall and becomes the spore.<br />

Variations of this pattern occur within<br />

the pro<strong>to</strong>stelids. For instance, some species<br />

produce spores which are discharged forcibly<br />

(e.g. Spiegel, 1984). In Ceratiomyxa fruticulosa,<br />

a species which may or may not belong <strong>to</strong><br />

the Pro<strong>to</strong>steliomycetes (Spiegel, 1990; Kirk et al.,<br />

2001; Clark et al., 2004), numerous spores are<br />

formed externally on a sporocarp (Figs. 2.7a,b)<br />

and are the product of meiosis. They germinate<br />

<strong>to</strong> release a single quadrinucleate pro<strong>to</strong>plast<br />

(Figs. 2.7c e) which divides repeatedly <strong>to</strong><br />

produce a clump of four and later eight haploid<br />

cells, the octette stage (Figs. 2.7f,g). Each of these<br />

cells releases a motile cell (a swarmer) which<br />

has one or two whiplash-type flagella (Fig. 2.7h).<br />

Fig 2.5 Pro<strong>to</strong>stelium sp. (a) Two sporocarps, one<br />

immature, the other with a detached spore. Note the<br />

apophysis beneath the spore. (b) Empty spore case after<br />

germination. (c) Amoeboid phase.


MYXOMYCETES: TRUE (PLASMODIAL) SLIME MOULDS<br />

47<br />

Fig 2.6 Sporogenesis in a<br />

pro<strong>to</strong>stelid (after Olive,1967).<br />

(a) Early pre-spore stage.<br />

(b) Hat-shaped stage.<br />

(c) Appearance of the steliogen.<br />

(d) Beginning of stalk formation.<br />

(e) Later stage in stalk development,<br />

with steliogen extending in<strong>to</strong><br />

upper part of stalk tube. (f) Mature<br />

sporocarp showing terminal spore,<br />

with subtending apophysis, outer<br />

sheath, and inner stalk tube.<br />

The swarmers eventually fuse <strong>to</strong> form a diploid<br />

zygote which initiates the plasmodial stage<br />

(Figs. 2.7i,j), from which the sporocarp develops<br />

(Spiegel, 1990). Ceratiomyxa fruticulosa thus<br />

shows features of both the Pro<strong>to</strong>steliomycetes<br />

in producing its spores externally, and the<br />

Myxomycetes (see below) in having a flagellated<br />

stage in its life cycle. Its precise phylogenetic<br />

position remains <strong>to</strong> be established. This species<br />

is probably homothallic (Clark et al., 2004).<br />

Its whitish semitransparent sporocarps are<br />

rather common on the surface of rotting wood<br />

(Plate 1a).<br />

2.5 Myxomycetes: true (plasmodial)<br />

slime moulds<br />

Fig 2.7 Ceratiomyxa fruticulosa. (a) Fruiting sporocarp bearing<br />

stalked spores. (b) Portion of the surface of the sporocarp<br />

showing spores and their attachment. (c) Spore. (d) Naked<br />

pro<strong>to</strong>plast emerging from the spore at germination. (e) Naked<br />

pro<strong>to</strong>plast before cleavage. (f) Cleavage of pro<strong>to</strong>plast <strong>to</strong> form<br />

a tetrad of pro<strong>to</strong>plasts. (g) Octette stage: a clump of eight<br />

pro<strong>to</strong>plasts. (h) Uniflagellate and biflagellate swarmer released<br />

from the octette pro<strong>to</strong>plasts. (i) Copulation of swarmers by<br />

their posterior ends. (j) Young plasmodium: c, contractile<br />

vacuole; s, ingested spore within food vacuole. (c i) <strong>to</strong> same<br />

scale.<br />

The Myxomycetes (zool.: Myxogastrea) are by<br />

far the largest group of slime moulds, comprising<br />

some 800 species in 62 genera which<br />

are currently divided in<strong>to</strong> five orders (Kirk<br />

et al., 2001). General accounts have been given<br />

by Frederick (1990), Stephenson and Stempen<br />

(1994) and Alexopoulos et al. (1996). A monograph<br />

of British species has been compiled by Ing<br />

(1999). These are the familiar slime moulds<br />

so common on moist, decaying wood and<br />

other organic substrata. They are also abundant<br />

in soil and may fulfil ecological functions<br />

which are as yet poorly unders<strong>to</strong>od (Madelin,<br />

1984).<br />

The vegetative phase is a free-living plasmodium,<br />

i.e. a multinucleate wall-less mass of<br />

pro<strong>to</strong>plasm. This may or may not be covered


48 PROTOZOA: MYXOMYCOTA (SLIME MOULDS)<br />

by a slime sheath. Plasmodia vary in size and can<br />

be loosely grouped in<strong>to</strong> three categories.<br />

(1) Pro<strong>to</strong>plasmodia are inconspicuous microscopic<br />

structures usually giving rise only <strong>to</strong> a<br />

single sporangium. They resemble the simple<br />

plasmodia of pro<strong>to</strong>stelids.<br />

(2) Aphanoplasmodia (Gr. aphanes ¼ invisible)<br />

are thin open networks of plasmodial strands.<br />

The aphanoplasmodium is transparent, with<br />

individual strands only 5 10 mm wide and the<br />

entire plasmodium about 100 200 mm in diameter.<br />

Most aphanoplasmodia are only seen<br />

with the aid of a dissection microscope.<br />

(3) Phaneroplasmodia (Gr. phaneros ¼ visible)<br />

are large sheets or networks with conspicuous<br />

veins (Fig. 2.8a) within which the pro<strong>to</strong>plasm<br />

shows rhythmic and reversible streaming, each<br />

pulse lasting about 60 90 s. This striking phenomenon<br />

is readily observed with a dissection<br />

microscope and is probably due <strong>to</strong> interactions<br />

of Ca 2þ ions with cy<strong>to</strong>skeletal elements lining<br />

the veins (see Section 2.5.3).<br />

2.5.1 Life cycle of myxomycetes<br />

The life cycle of Physarum polycephalum, a typical<br />

myxomycete, is summarized in Fig. 2.9. The plasmodium<br />

is diploid and feeds by phagocy<strong>to</strong>sis of<br />

bacteria, yeasts or fungal mycelia or spores.<br />

It gives rise <strong>to</strong> a sporophore under appropriate<br />

conditions. The haploid spores are dispersed<br />

Fig 2.8 Phaneroplasmodia of Physarum polycephalum.<br />

(a) Margin of extending plasmodium.The pro<strong>to</strong>plasm<br />

is particularly dense at the advancing edge.Further<br />

behind, pro<strong>to</strong>plasm is concentrated in large veins<br />

which show rhythmic pulsation. (b) Fusion between<br />

compatible plasmodia. Note the complete fusion of<br />

veins. (c) Lethal reaction following fusion between<br />

incompatible plasmodia. (a) from Carlile (1971),<br />

(b) and (c) from Carlile and Dee (1967), by permission<br />

of Academic Press (a) and Macmillan Journals (b,c).<br />

Original prints kindly supplied by M. J.Carlile.


MYXOMYCETES: TRUE (PLASMODIAL) SLIME MOULDS<br />

49<br />

Fig 2.9 Life cycle of the myxomycete Physarum polycephalum. Spores released from the sporangium are haploid and can germinate<br />

by releasing either a single myxamoeba or a swarmer cell.These two cell types are interconvertible.The myxamoeba can divide<br />

mi<strong>to</strong>tically. In P. polycephalum, plasmogamy (P) usually takes place between swarmers which must belong <strong>to</strong> different mating types.<br />

Karyogamy (K) follows, and the diploid zygote establishes a phaneroplasmodium.When nutrients become limiting, a sporophore is<br />

formed and differentiates sporangia in which meiosis (M) occurs.Unfavourable conditions can be overcome at the haploid stage<br />

when the myxamoeba forms a microcyst, or at the diploid stage when the plasmodium forms sclerotia.Open and closed circles<br />

represent haploid nuclei of opposite mating type; diploid nuclei are larger and half-filled.<br />

by wind or insects and, depending on environmental<br />

conditions such as moisture, germinate<br />

by releasing either amoebae or zoospores<br />

(swarmers) with usually two anterior whiplash<br />

flagella, of which one is shorter than the other<br />

and is thus often invisible (Fig. 2.10). The amoebae<br />

are called myxamoebae, in order <strong>to</strong> distinguish<br />

them from the amoebae of cellular<br />

slime moulds which have a different function<br />

in the life cycle. Myxamoebae are capable of<br />

asexual reproduction by division. Swarmers<br />

cannot divide, but can readily and reversibly<br />

convert in<strong>to</strong> myxamoebae. Under adverse conditions,<br />

myxamoebae secrete a wall <strong>to</strong> form<br />

microcysts. Both swarmers and myxamoebae<br />

form filose pseudopodia with which they engulf<br />

their prey. Sexual reproduction is initiated when<br />

two haploid myxamoebae or swarmers of compatible<br />

mating type fuse <strong>to</strong> form a zygote from<br />

which the diploid plasmodium develops. The<br />

plasmodium can survive adverse conditions by<br />

turning in<strong>to</strong> a resistant sclerotium in which<br />

numerous walled compartments (spherules),<br />

each containing several nuclei, are formed.<br />

Upon resumption of growth, the pro<strong>to</strong>plasts<br />

emerge from their spherules and fuse <strong>to</strong> re-establish<br />

the plasmodium. When sexual reproduction<br />

ensues, the entire content of a plasmodium is<br />

converted in<strong>to</strong> one or more sporangia in which<br />

meiosis takes place. Beneath the developing sporangia,<br />

the plasmodium deposits a specialized<br />

layer, the hypothallus, which is very variable in


50 PROTOZOA: MYXOMYCOTA (SLIME MOULDS)<br />

Fig 2.10 Spore germination and swarmers in Physarum and Reticularia.a.Physarum polycephalum:1,sporesgerminating <strong>to</strong>release<br />

myxamoebae; 2, uniflagellate and biflagellate swarmers, note the pseudopodia at the front end of one swarmer; 3, myxamoeba;<br />

4, fusion between two myxamoebae. b. Reticularia lycoperdon:1, spore showing cracked wall; 2, swarmers, one with pseudopodia;<br />

3, encystment stage; 4, fusion between two swarmers.<br />

form (disc-like, membranous, horny or spongy).<br />

In P. polycephalum, sexual reproduction is triggered<br />

by environmental fac<strong>to</strong>rs such as<br />

starvation and light, and by chemical fac<strong>to</strong>rs,<br />

e.g. Ca 2þ and malate (Renzel et al., 2000).<br />

Depending on species, the sporophores may<br />

take a range of shapes. Intermediates between<br />

these different types of sporophore are possible.<br />

The most common form is the sporangium,<br />

a vessel enclosed by a wall (peridium) within<br />

which the spores are contained (Plates 1e,f,h).<br />

Pro<strong>to</strong>plasmodia produce only one sporangium<br />

each, but numerous sporangia may arise from<br />

phaneroplasmodia. Sporangia may be stalked<br />

or sessile. A second common sporophore is the<br />

aethalium (Gr. aethes ¼ irregular, curious,<br />

unusual) in which the entire plasmodium<br />

becomes converted in<strong>to</strong> a hemispherical or<br />

cushion-shaped structure (Plates 1b d,g). This<br />

can comprise several sporangia, but these have<br />

usually lost their structural identity and are<br />

surrounded by one common peridium. In a<br />

pseudoaethalium, several sporangia are grouped<br />

<strong>to</strong>gether but are still recognized as structurally<br />

distinct. In the plasmodiocarp, the pro<strong>to</strong>plasm<br />

accumulates in the main veins of the plasmodium,<br />

and spores are produced there.<br />

Frederick (1990) has described methods for<br />

the isolation and cultivation of myxomycetes.<br />

Some species, such as Physarum polycephalum,<br />

can be grown in axenic culture and have<br />

become valuable systems for experimentation.<br />

Other species need <strong>to</strong> be fed with bacteria or<br />

sterile oat flakes. Plasmodia can be maintained<br />

for prolonged periods in a vegetative state, and<br />

sclerotia can be s<strong>to</strong>red dry for months. Spores


MYXOMYCETES: TRUE (PLASMODIAL) SLIME MOULDS<br />

51<br />

have been revived after more than 50 years’<br />

s<strong>to</strong>rage in a herbarium (Elliott, 1949).<br />

2.5.2 Orders of myxomycetes<br />

Myxomycetes are currently grouped in<strong>to</strong> five<br />

orders, all of which are frequently found either<br />

in nature or upon incubating suitable plant<br />

material on moist filter paper.<br />

The Echinosteliales (e.g. Echinostelium,<br />

Clas<strong>to</strong>derma) contain the smallest known<br />

myxomycetes. They form pro<strong>to</strong>plasmodia, with<br />

each pro<strong>to</strong>plasmodium giving rise <strong>to</strong> only one<br />

sporangium. The Echinosteliales resemble the<br />

pro<strong>to</strong>stelids from which they are probably<br />

derived (Frederick, 1990; Spiegel, 1991; see<br />

Fig. 2.5).<br />

The Liceales (e.g. Lycogala, Dictydium, Cribraria,<br />

Reticularia) are common on the bark of dead<br />

trees. Some of the smaller species produce pro<strong>to</strong>plasmodia,<br />

but most have phaneroplasmodia.<br />

Various types of sporophores are formed; the<br />

aethalia of Lycogala epidendron (Plate 1b) and<br />

Reticularia (= Enteridum) lycoperdon (Plates 1c,d)<br />

are particularly common.<br />

The Trichiales (e.g. Arcyria, Trichia, Hemitrichia)<br />

are ubiqui<strong>to</strong>us on fallen logs. The plasmodia are<br />

intermediate between aphanoplasmodia and<br />

phaneroplasmodia. Fructifications in Trichia floriforme<br />

are well-defined sporangia which contain<br />

an internal meshwork of threads, collectively<br />

called the capillitium. The peridium breaks<br />

open at maturity, and the spores are released<br />

over time by the twisting of the capillitial<br />

threads which thus act as elaters (Fig. 2.11).<br />

Arcyria denudata produces reddish sporangia<br />

on rotting wood (Plate 1e). Another member,<br />

Hemitrichia serpula, produces plasmodiocarps.<br />

The Physarales (e.g. Physarum, Fuligo) produce<br />

the largest plasmodia. Physarum polycephalum has<br />

been used extensively in fundamental research<br />

on cell biology, for example on the nature<br />

of pro<strong>to</strong>plasmic streaming, or the synchrony of<br />

nuclear division in a large plasmodium comprising<br />

thousands of nuclei (see below). The plasmodia<br />

are typical phaneroplasmodia, each of<br />

which produces numerous sporangia at maturity<br />

(Plate 1f). Fuligo septica forms particularly large<br />

sporophores (aethalia) which are bright yellow<br />

and are frequently seen on decaying wood<br />

(Plate 1g).<br />

The Stemonitales include such genera as<br />

Comatricha and Stemonitis. Stemonitis spp. produce<br />

clusters of stalked sporangia from aphanoplasmodia<br />

which are visible on rotting wood<br />

(Plate 1h).<br />

2.5.3 Physarum polycephalum as an<br />

experimental <strong>to</strong>ol<br />

This species has been used <strong>to</strong> investigate several<br />

aspects of cell biology. The conspicuous cy<strong>to</strong>plasmic<br />

shuttle streaming in the veins of its large<br />

phaneroplasmodia is a fascinating phenomenon<br />

and has been examined extensively. The pulse is<br />

caused by actin myosin interactions controlled<br />

by Ca 2þ (Smith, 1994). It is brought about not<br />

by the direct binding of organelles <strong>to</strong> actin<br />

cables, but by the constriction and relaxation<br />

of an actin myosin skele<strong>to</strong>n lining the veins.<br />

Several proteins interacting with actin and<br />

myosin are directly or indirectly regulated by<br />

Ca 2þ , but the most important effect of Ca 2þ<br />

is on one of the myosin light chains. This is a<br />

regula<strong>to</strong>ry subunit which directly binds Ca 2þ .<br />

In contrast <strong>to</strong> most animal actin myosin<br />

systems which are stimulated by Ca 2þ , that of<br />

Physarum is inhibited, i.e. contraction occurs<br />

at low Ca 2þ concentrations, and relaxation at<br />

higher concentrations. Ca 2þ -inhibited actin<br />

myosin interaction also occur in plant cells<br />

where they are visible as cy<strong>to</strong>plasmic streaming.<br />

Nakamura and Kohama (1999) have written a<br />

thorough review of the actin myosin system<br />

in Physarum.<br />

Mi<strong>to</strong>tic division of all nuclei throughout<br />

the plasmodium of P. polycaphalum occurs in a<br />

synchronized manner, and Physarum was one<br />

of the pioneer organisms in which the existence<br />

of the cell cycle was demonstrated. Synchrony of<br />

mi<strong>to</strong>sis is regulated by a protein kinase which<br />

catalyses the phosphorylation of H1 his<strong>to</strong>nes,<br />

leading <strong>to</strong> the condensation of chromosomes at<br />

the onset of mi<strong>to</strong>sis (Bradbury et al., 1974; Inglis<br />

et al., 1976). This protein kinase is now known <strong>to</strong><br />

be homologous <strong>to</strong> the cdc2 product in the fission<br />

yeast Schizosaccharomyces pombe (see Fig. 9.5;<br />

Langan et al., 1989).


52 PROTOZOA: MYXOMYCOTA (SLIME MOULDS)<br />

Fig 2.11 Trichia floriforme. (a) Undehisced<br />

sporangia. Note that the sporangial stalks are<br />

continuous with the hypothallus. (b) Dehisced<br />

sporangia releasing spores by twisting of<br />

elaters. (c) Elaters and spores.<br />

A further interesting feature of P. polycephalum<br />

is the behaviour of the plasmodium<br />

and the manner in which its actions are coordinated.<br />

Little work has been carried out<br />

beyond descriptions of striking phenomena.<br />

One is the ability of P. polycephalum plasmodia<br />

<strong>to</strong> find the shortest way <strong>to</strong> a food source through<br />

an artificially constructed maze (Nakagaki, 2001).<br />

Another is the pattern of veins which is established<br />

when different regions of a plasmodium<br />

are presented with food sources; the configuration<br />

of the plasmodium has been called a ‘smart<br />

network’ because it presents the shortest<br />

possible <strong>to</strong>tal length of veins <strong>to</strong> provide good<br />

interconnections while making allowances for<br />

blockage of individual veins (Nakagaki et al.,<br />

2004).<br />

When separate plasmodia of P. polycephalum<br />

or other species meet, two reactions are possible,<br />

i.e. a compatible reaction in which the plasmodia<br />

fuse and their veins coalesce (Fig. 2.8b) or an<br />

incompatible reaction in which the plasmodia<br />

fail <strong>to</strong> fuse and move away from each other, or<br />

fusion is attempted but stalls and is followed by<br />

death of the fusion regions of both plasmodia<br />

(Fig. 2.8c). This is called the lethal reaction.


MYXOMYCETES: TRUE (PLASMODIAL) SLIME MOULDS<br />

53<br />

Genetic studies have shown that fusion occurs<br />

between plasmodia of genetically closely related<br />

strains (Carlile & Dee, 1967). The type of incompatibility<br />

brought about by the interaction of<br />

genetically distinct plasmodia is an example<br />

of a widespread phenomenon called vegetative<br />

incompatibility which is found not only in<br />

slime moulds, but also in the Eumycota, vertebrates<br />

and other organisms. In humans, a similar<br />

phenomenon accounts for blood grouping or<br />

the failure of tissue transplantations. It is interesting<br />

<strong>to</strong> consider the paradox that fusion<br />

between genetically dissimilar myxamoebae is<br />

encouraged during sexual reproduction by the<br />

existence of different mating types, whereas<br />

it is discouraged during vegetative fusion of<br />

plasmodia.


3<br />

Pro<strong>to</strong>zoa: Plasmodiophoromycota<br />

3.1 <strong>Introduction</strong><br />

The Plasmodiophoromycota are a group of<br />

obligate (i.e. biotrophic) parasites. The bestknown<br />

examples attack higher plants, causing<br />

economically significant diseases such as clubroot<br />

of brassicas (Plasmodiophora brassicae),<br />

powdery scab of pota<strong>to</strong> (Spongospora subterranea;<br />

formerly S. subterranea f. sp. subterranea) and<br />

crook-root disease of watercress (S. nasturtii; formerly<br />

S. subterranea f. sp. nasturtii). In addition <strong>to</strong><br />

damaging crops directly, some species (S. subterranea,<br />

Polymyxa betae, P. graminis) also act as<br />

vec<strong>to</strong>rs for important plant viruses (Adams, 1991;<br />

Campbell, 1996). Other species infect roots<br />

and shoots of non-cultivated plants, especially<br />

aquatic plants. Algae, dia<strong>to</strong>ms and Oomycota are<br />

also attacked. If the nine species of Hap<strong>to</strong>glossa,<br />

which parasitize nema<strong>to</strong>des and rotifers, are<br />

included in the Plasmodiophoromycota, the<br />

phylum currently comprises 12 genera and 51<br />

species (Dick, 2001a). Genera are separated<br />

from each other largely by the arrangement of<br />

resting spores in the host cell (Waterhouse,<br />

1973). This feature has also been used for<br />

naming most genera; for instance, in Polymyxa,<br />

numerous resting spores are contained within<br />

each sorus, whereas in Spongospora the resting<br />

spores are grouped loosely in a sponge-like<br />

sorus (Fig. 3.6). Accounts of the Plasmodiophoromycota<br />

have been given by Sparrow (1960),<br />

Karling (1968), Dylewski (1990) and Brasel<strong>to</strong>n<br />

(1995, 2001).<br />

3.1.1 Taxonomic considerations<br />

Plasmodiophoromycota have traditionally been<br />

studied by mycologists and plant pathologists.<br />

Many general features of their biology and<br />

epidemiology are similar <strong>to</strong> those of certain<br />

members of the Chytridiomycota such as<br />

Olpidium (see p. 145). However, it is now clear<br />

from DNA sequence analysis and other criteria<br />

that Plasmodiophora is related neither <strong>to</strong> the<br />

Oomycota and other Straminipila (Chapters 4<br />

and 5) nor <strong>to</strong> the true fungi (Eumycota). Instead,<br />

it is distantly related <strong>to</strong> the Myxomycota<br />

discussed in Chapter 2 but belongs <strong>to</strong> a different<br />

grouping within the Pro<strong>to</strong>zoa (Barr, 1992;<br />

Castlebury & Domier, 1998; Ward & Adams,<br />

1998; Archibald & Keeling, 2004).<br />

Some believe that Hap<strong>to</strong>glossa is related <strong>to</strong> the<br />

Oomycota rather than Pro<strong>to</strong>zoa, although no<br />

molecular data seem <strong>to</strong> be available as yet <strong>to</strong><br />

support this claim. Since Hap<strong>to</strong>glossa strikingly<br />

resembles Plasmodiophora in its infection biology,<br />

we shall include it in this chapter. With the<br />

possible exception of Hap<strong>to</strong>glossa, the phylum<br />

Plasmodiophoromycota is monophyletic and contains<br />

a single class (Plasmodiophoromycetes). We<br />

consider two orders in this chapter, Plasmodiophorales<br />

and Hap<strong>to</strong>glossales.<br />

3.2 Plasmodiophorales<br />

The zoospore of the Plasmodiophorales is biflagellate.<br />

The flagella are inserted laterally and are


PLASMODIOPHORALES<br />

55<br />

of unequal length, the anterior one being<br />

shorter. Both flagella are of the whiplash type<br />

(Fig. 1.17c). Zoospores of this type are said <strong>to</strong> be<br />

anisokont. Transmission electron microscopy<br />

(TEM) studies have shown that the tips of the<br />

flagella are tapered rather than blunt (Clay &<br />

Walsh, 1997). Like the zoospore, the main<br />

vegetative unit the amoeba, which enlarges<br />

<strong>to</strong> become a plasmodium is wall-less. It is<br />

present freely within host plant cells, its<br />

membrane being in direct contact with the<br />

host cy<strong>to</strong>plasm. The plasmodia possess amoeboid<br />

features because they can produce pseudopodia<br />

and engulf parts of the host cy<strong>to</strong>plasm by<br />

phagocy<strong>to</strong>sis (Clax<strong>to</strong>n et al., 1996; Clay & Walsh,<br />

1997). This has been interpreted as a primitive<br />

trait perhaps betraying a free-living amoeboid<br />

ances<strong>to</strong>r with a phagocy<strong>to</strong>tic mode of nutrition<br />

(Buczacki, 1983). Some Plasmodiophorales can<br />

now be grown away from their host on artificial<br />

media for prolonged periods if bacteria are<br />

present. These are phagocy<strong>to</strong>sed by amoeboid<br />

growth forms (Arnold et al., 1996). In their hosts,<br />

amoeboid plasmodia can digest their way<br />

through plant cell walls, moving <strong>to</strong> adjacent<br />

uninfected cells and thus spreading the infection<br />

within an infected root (Mithen & Magrath, 1992;<br />

Clax<strong>to</strong>n et al., 1996).<br />

The walled stages of Plasmodiophorales are<br />

confined <strong>to</strong> the zoospore cysts on the plant<br />

surface, and the zoosporangia and resting sporangia<br />

inside host plant cells. The wall of resting<br />

spores is particularly thick and has been shown<br />

<strong>to</strong> contain chitin (Moxham & Buczacki, 1983).<br />

3.2.1 Life cycle of Plasmodiophorales<br />

Certain details of the life cycle of the<br />

Plasmodiophorales are still doubtful (Fig. 3.1).<br />

However, the known stages show very little<br />

variation between different species, indicating<br />

that the life cycle is conserved throughout the<br />

order. A resting spore germinates by releasing<br />

a single haploid zoospore (primary zoospore)<br />

which encysts on a suitable surface by secreting a<br />

cell wall. After a while, an amoeba is injected<br />

from the cyst in<strong>to</strong> a host cell such as a root hair<br />

where it enlarges <strong>to</strong> form a plasmodium, accompanied<br />

by mi<strong>to</strong>tic nuclear divisions. Nuclear<br />

divisions at this stage are cruciform; the nucleolus<br />

is prominently visible throughout the mi<strong>to</strong>tic<br />

process, elongating in two directions <strong>to</strong> take<br />

up a cross-like shape when viewed in certain<br />

sections by transmission electron microscopy.<br />

This feature is unique <strong>to</strong> the Plasmodiophorales<br />

(Brasel<strong>to</strong>n, 2001). After a while, nuclei divide<br />

mi<strong>to</strong>tically in a non-cruciform manner, and the<br />

contents of the plasmodium differentiate in<strong>to</strong><br />

zoospores. This type of plasmodium is termed<br />

the primary plasmodium or sporangial plasmodium<br />

because it produces zoospores. The zoospores<br />

are called secondary zoospores because<br />

they arise from a sporangium, not from a resting<br />

spore. Once released, secondary zoospores may<br />

re-infect the host <strong>to</strong> give rise <strong>to</strong> further primary<br />

plasmodia and zoosporangia. Eventually, however,<br />

a different type of plasmodium, the secondary<br />

plasmodium or sporogenic plasmodium, is<br />

formed which undergoes meiotic nuclear divisions<br />

and produces resting spores (Garber & Aist,<br />

1979; Brasel<strong>to</strong>n, 1995). It is not known where<br />

plasmogamy and karyogamy occur in the life<br />

cycle of the Plasmodiophorales.<br />

All developmental stages of P. brassicae can be<br />

produced readily in the labora<strong>to</strong>ry. Clubbed<br />

roots should be collected from a field or garden<br />

and kept frozen at 20°C. Seedlings of brassicas,<br />

susceptible Chinese cabbage cultivars or<br />

Arabidopsis thaliana should be grown in a soil<br />

with a high peat content which must be kept<br />

well watered. Infections can be established<br />

by adding slices of infected root material<br />

or a resting spore suspension <strong>to</strong> the soil.<br />

Zoosporangia will be formed within a few days,<br />

and root galls should be visible within 3 7 weeks<br />

(Castlebury & Glawe, 1993). Pota<strong>to</strong> or <strong>to</strong>ma<strong>to</strong><br />

plants can be infected with Spongospora subterranea<br />

using similar pro<strong>to</strong>cols. Cabbage callus<br />

cultures are occasionally used as a simplified<br />

experimental system for life cycle studies of<br />

P. brassicae (Tommerup & Ingram, 1971).<br />

3.2.2 Plasmodiophora brassicae<br />

Plasmodiophora brassicae is the causal organism of<br />

club root or finger-and-<strong>to</strong>e disease of brassicas<br />

(Fig. 3.2) and was first described by Woronin<br />

(1878). The disease is common in gardens where


56 PROTOZOA: PLASMODIOPHOROMYCOTA<br />

Fig 3.1 Probable life cycle of Plasmodiophora brassicae. A haploid resting spore forms a haploid primary zoospore giving rise <strong>to</strong> a<br />

multinucleate haploid primary plasmodium upon infection of a root hair. Secondary zoospores are also haploid, and the way in<br />

which they meet <strong>to</strong> form a secondary heterokaryotic plasmodium is not known for sure.Open and closed circles represent haploid<br />

nuclei of opposite mating type; the position of the diploidphase in the life cycle is unclear.Key events in the life cycle are plasmogamy<br />

(P), karyogamy (K) and meiosis (M). AfterTommerup and Ingram (1971), Buczacki (1983) and Dylewski (1990).<br />

brassicas are frequently grown, especially if the<br />

soil is acidic and poorly drained. A wide range of<br />

brassicaceous hosts is attacked, and root-hair<br />

infection of some non-brassicaceous hosts can<br />

also occur (Ludwig-Müller et al., 1999). The<br />

disease is widely distributed throughout the<br />

world.<br />

Club root symp<strong>to</strong>ms<br />

Infected crucifers usually have greatly swollen<br />

roots. Both tap roots and lateral roots may be<br />

affected. Occasionally, infection results in the<br />

formation of adventitious root buds which give<br />

rise <strong>to</strong> swollen stunted shoots. Above ground,<br />

however, infected plants may be difficult <strong>to</strong><br />

distinguish from healthy ones. The first symp<strong>to</strong>m<br />

is wilting of the leaves in warm weather,<br />

although such wilted leaves often recover at<br />

night. Later the rate of growth of infected plants<br />

is retarded so that they appear yellow and<br />

stunted. Plants infected at the seedling stage<br />

may be killed, but if infection is delayed the<br />

effect is much less severe and well-developed<br />

heads of cabbage, cauliflower, etc., can form on<br />

plants with quite extensive root hypertrophy<br />

(swelling of cells) and hyperplasia (enhanced


PLASMODIOPHORALES<br />

57<br />

Fig 3.2 Club root of cabbage caused by<br />

Plasmodiophora brassicae.<br />

division of cells). Microscopically, even infected<br />

root hairs are expanded at their tips <strong>to</strong> form<br />

club-shaped swellings which are sometimes<br />

lobed and branched (Fig. 3.3). Rausch et al.<br />

(1981) followed the growth of infected and<br />

uninfected seedlings of Chinese cabbage, a<br />

particularly susceptible host. Within the first 30<br />

days, the growth rates of infected and control<br />

plants were almost identical, and clubs developed<br />

in proportion <strong>to</strong> shoot growth. Wilting of<br />

infected plants was observed beyond 30 days<br />

when the clubs developed at the expense of<br />

shoots. Plants growing in suboptimal conditions,<br />

e.g. in the shade, produced disproportionately<br />

smaller clubs. Generally, the root/shoot ratio is<br />

appreciably higher in infected plants, suggesting<br />

a diversion of pho<strong>to</strong>synthetic product <strong>to</strong> the<br />

clubbed roots. The P. brassicae infection therefore<br />

acts as a new carbon sink.<br />

The process of infection<br />

Swollen roots contain a large number of small<br />

spherical resting spores, and when these roots<br />

decay the spores are released in<strong>to</strong> the soil.<br />

Electron micrographs show that the resting<br />

spores have spiny walls (Yukawa & Tanaka,<br />

1979). The resting spore germinates <strong>to</strong> produce<br />

a single zoospore with two flagella of unequal<br />

length, both of the whiplash type and with the<br />

usual 9 þ 2 arrangement of microtubules (Aist &<br />

Williams, 1971). Germination of resting spores is<br />

stimulated by substances specific <strong>to</strong> Brassicaceae,<br />

possibly allyl isothiocyanates, which diffuse from<br />

the cabbage roots in<strong>to</strong> the soil (Macfarlane,<br />

1970).<br />

The primary zoospore (i.e. the first motile<br />

stage released from the resting spore) swims by<br />

means of its flagella, the long flagellum trailing<br />

and the short one pointing forward. The process<br />

of root hair infection has been followed in a<br />

classical study by Aist and Williams (1971). Since<br />

the first such study, on penetration by Polymyxa<br />

betae, was written in German (Keskin & Fuchs,<br />

1969), the German terminology is still in use<br />

<strong>to</strong>day. Primary zoospores of P. brassicae are<br />

released some 26 30 h after placing a suspension<br />

of resting spores close <strong>to</strong> seedling roots of<br />

cabbage. The zoospores may collide several times<br />

with a root hair before becoming attached, and<br />

appear <strong>to</strong> be attached at a point opposite <strong>to</strong> the<br />

origin of the flagella.<br />

The flagella coil around the zoospore body,<br />

which becomes flattened against the host wall,<br />

and pseudopodium-like extensions of the zoospore<br />

develop, being continuously extended and<br />

withdrawn. The flagella are then withdrawn, and<br />

the zoospore encysts, attached <strong>to</strong> the root hair<br />

(Fig. 3.4). The zoospore cyst contains lipid bodies<br />

and a vacuole which enlarges during cyst<br />

maturation, which takes a few hours. The most<br />

conspicuous ultrastructural feature of mature<br />

cysts is a long Rohr (tube), with its outer end<br />

pointing <strong>to</strong>wards the root hair wall. This end<br />

of the tube is occluded by a plug. Within the tube


58 PROTOZOA: PLASMODIOPHOROMYCOTA<br />

Fig 3.3 Plasmodiophora brassicae.(a)T.S.<br />

through young infected cabbage root showing<br />

secondary (sporogenic) plasmodia in the<br />

cortex. Note the hypertrophy of some of the<br />

host cells containing plasmodia, and the<br />

presence of young plasmodia in cells<br />

immediately outside the xylem. (b) T.S. cabbage<br />

root at a later stage of infection, showing the<br />

formation of resting spores. (c) Primary<br />

(zoosporangial) plasmodium in cabbage root<br />

hair 4 days after planting in a heavily<br />

contaminated soil. (d) Young primary<br />

zoosporangia in root hair. Note the<br />

club-shaped swelling of the infected root hair.<br />

(e) Mature and discharged primary<br />

zoosporangia. a and b <strong>to</strong> same scale; (c e) <strong>to</strong><br />

same scale.<br />

Fig 3.4 Plasmodiophora brassicae. (a) Resting<br />

spores, one full, one empty (showing a pore in<br />

the wall). (b) Zoospore. (c) Attachment of<br />

zoospore <strong>to</strong> root hair. (d) Zoospore cyst with<br />

adhesorium following withdrawal of flagellar<br />

axonemes. (e) Entry of amoeba in<strong>to</strong> root hair.<br />

Based on Aist and Williams (1971).


PLASMODIOPHORALES<br />

59<br />

Fig 3.5 Plasmodiophora brassicae. Diagrammatic summary of penetration process (after Aist & Williams,1971).The diagram shows<br />

a zoospore cyst attached <strong>to</strong> the wall of a root hair. (a) Cyst vacuole not yet enlarged. (b) About 3 h later, the cyst vacuole enlarges<br />

and a small adhesorium appears. (c) About1min later, the stylet punctures the host cell wall. (d) Penetration has occurred and the<br />

host pro<strong>to</strong>plast has deposited a papilla at the penetration site.<br />

is a bullet-shaped Stachel (stylet), the outer part<br />

of which is made up of parallel fibrils. Behind the<br />

blunt posterior end of the stylet, the tube<br />

narrows <strong>to</strong> form a Schlauch (sac).<br />

Penetration of the root hair wall occurs about<br />

3 h after encystment, as after this time the first<br />

empty vacuolated cysts are observed. The penetration<br />

process takes place rapidly, and an<br />

interpretation of it is shown in Fig. 3.5. Firm<br />

attachment of the tube <strong>to</strong> the root hair is<br />

brought about by the adhesorium, which may<br />

develop by partial evagination (i.e. turning inside<br />

out) of the tube (Fig. 3.5b). During evagination,<br />

an adhesive substance which has a fibrillar<br />

appearance in TEM micrographs is released<br />

on<strong>to</strong> the adhesorial surface from its s<strong>to</strong>rage site<br />

inside the tube. The enlargement of the vacuole<br />

is presumably the driving force which brings<br />

about complete evagination of the tube within<br />

1 min, followed by thrusting the stylet through<br />

the host wall. The pathogen is injected in<strong>to</strong> the<br />

host cell as a small, spherical, wall-less amoeba<br />

which becomes caught up by cy<strong>to</strong>plasmic streaming.<br />

After penetration (Fig. 3.5d), a papilla of<br />

callose is deposited around the penetration<br />

point beneath the adhesorium, possibly as a<br />

wound-healing response. Similar penetration<br />

mechanisms have been described for other<br />

Plasmodiophorales, including Spongospora subterranea<br />

(Merz, 1997), S. nasturtii (Clax<strong>to</strong>n et al., 1996)<br />

and Polymyxa betae (Keskin & Fuchs, 1969). Details<br />

of the infection process by P. betae have been<br />

filmed (see Webster, 2006a). A yet more elaborate<br />

process of infection is found in Hap<strong>to</strong>glossa,<br />

which parasitizes nema<strong>to</strong>des and rotifers<br />

(see p. 65).<br />

Development of zoosporangia<br />

Within the infected root hair, the amoeba may<br />

divide in<strong>to</strong> several uninucleate amoebae. Later<br />

the nuclei within each amoeba show cruciform<br />

divisions, giving rise <strong>to</strong> small multinucleate


60 PROTOZOA: PLASMODIOPHOROMYCOTA<br />

primary plasmodia. Each plasmodium divides up<br />

<strong>to</strong> form a group (sorus) of roughly spherical<br />

thin-walled zoosporangia lying packed <strong>to</strong>gether<br />

in the host cell (Fig. 3.3). Separate pro<strong>to</strong>plasts<br />

might coalesce at this stage. Each zoosporangium<br />

finally contains 4 8 uninucleate zoospores.<br />

These are morphologically identical <strong>to</strong><br />

primary zoospores. Some mature zoosporangia<br />

become attached <strong>to</strong> the host cell wall and an exit<br />

pore develops at this point through which the<br />

zoospores escape. The zoospores of other sporangia<br />

are released in<strong>to</strong> those with an exit pore.<br />

Occasionally, zoospores escape in<strong>to</strong> the lumen of<br />

the host cell. Liberated zoospores can re-infect<br />

plant roots, thereby completing an asexual cycle<br />

(Fig. 3.1).<br />

Sexual reproduction<br />

In P. brassicae, resting sporangia are not formed<br />

in root hairs after the first cycle of infection, but<br />

are located mainly in older infections in strongly<br />

hypertrophied regions of the root cortex. There is<br />

evidence that resting sporangia are involved in<br />

sexual reproduction (Fig. 3.1) because meiotic<br />

nuclear divisions with synap<strong>to</strong>nemal complexes<br />

have been observed in maturing resting sporangia<br />

(Garber & Aist, 1979). Further, each resting<br />

spore normally contains one haploid nucleus<br />

(Narisawa et al., 1996). <strong>Third</strong>ly, infection experiments<br />

have established that resting sporangia<br />

are formed only if two genetically dissimilar<br />

nuclei are present (Narisawa & Hashiba, 1998)<br />

which could be contributed either by two<br />

uninucleate zoospores or by a binucleate<br />

zoospore.<br />

The positions of the preceding stages of<br />

sexual reproduction plasmogamy and karyogamy<br />

in the life cycle of P. brassicae are still a<br />

matter of doubt. One possibility is that secondary<br />

zoospores fuse <strong>to</strong> form a dikaryon, followed by<br />

karyogamy. Quadriflagellate binucleate swarmers<br />

have indeed been observed and can result<br />

from the fusion of zoospores (Tommerup &<br />

Ingram, 1971). However, it is not yet clear<br />

whether these quadriflagellate spores can infect<br />

plant cells from the outside. Quadriflagellate<br />

binucleate zoospores may also arise from incomplete<br />

cleavage of cy<strong>to</strong>plasm during zoospore<br />

formation.<br />

Plasmodia of P. brassicae have been shown <strong>to</strong><br />

break through plant cell walls, thereby spreading<br />

an infection from root hairs in<strong>to</strong> deeper tissues<br />

of the root cortex (Mithen & Magrath, 1992).<br />

A conceivable alternative would be their migration<br />

through plasmodesmata. It is possible that<br />

two primary plasmodia or uninucleate amoebae<br />

arising from separate root hair infections fuse<br />

upon encountering each other deep inside the<br />

host plant. Such a fusion would produce a<br />

secondary plasmodium, and could be followed<br />

by karyogamy and meiosis, which would lead <strong>to</strong><br />

the development of resting spores (Fig. 3.1).<br />

Hypertrophy of infected host cells<br />

As the plasmodia within a host cell<br />

enlarge, the host nucleus remains active and<br />

undergoes repeated divisions. Hypertrophy and<br />

an increased ploidy of the host nuclei result, at<br />

least in callus culture experiments, because the<br />

mechanism for host cell division is apparently<br />

blocked (Tommerup & Ingram, 1971).<br />

Unsurprisingly, the grossly hypertrophied<br />

clubs contain enhanced levels of plant growth<br />

hormones. The concentration of auxins (especially<br />

indole-3-acetic acid, IAA) in clubbed roots<br />

was measured <strong>to</strong> be about 1.7 times as high as in<br />

uninfected roots (Ludwig-Müller et al., 1993), and<br />

that of cy<strong>to</strong>kinins was 2 3 times elevated<br />

(Dekhuijzen, 1980). Isolated secondary plasmodia<br />

of P. brassicae have been demonstrated <strong>to</strong> synthesize<br />

the cy<strong>to</strong>kinin zeatin (Müller & Hilgenberg,<br />

1986), and the amount of zeatin produced would<br />

be sufficient <strong>to</strong> establish a new carbon sink. The<br />

situation is more complicated with respect <strong>to</strong><br />

auxins which are not synthesized by plasmodia.<br />

Instead, the pathogen interferes with the host’s<br />

auxin metabolism, which is complex (Normanly,<br />

1997). The tissues of healthy crucifers contain<br />

relatively large amounts of indole glucosinolates<br />

such as glucobrassicin (¼ indole-3-methylglucosinolate)<br />

which is converted by the enzyme<br />

myrosinase <strong>to</strong> 3-indoleace<strong>to</strong>nitrile (IAN), a<br />

direct IAA precursor. Conversion of IAN <strong>to</strong> IAA<br />

is catalysed by nitrilase. Increased concentrations<br />

of indole glucosinolates, IAN and IAA have<br />

been measured in clubbed roots (Ludwig-Müller,<br />

1999), and the expression of nitrilase and<br />

myrosinase was also enhanced. Further, nitrilase


PLASMODIOPHORALES<br />

61<br />

protein was detectable by immunohis<strong>to</strong>chemical<br />

methods only in cells containing sporulating<br />

plasmodia. The activities of the above enzymes<br />

might be regulated by the signalling molecule,<br />

jasmonic acid (Grsic et al., 1999). However, these<br />

metabolic changes were confined <strong>to</strong> a narrow<br />

window of time, and other sources of IAA, such<br />

as its release from IAA alanine conjugates by<br />

the activity of amidohydrolase, are likely <strong>to</strong><br />

contribute (Ludwig-Müller et al., 1996). The<br />

host pathogen interactions leading <strong>to</strong> enhanced<br />

auxin levels in clubbed roots are therefore very<br />

intricate.<br />

At first, only cortical cells of the young root<br />

are infected, but later small plasmodia can be<br />

found in the medullary ray cells and in the<br />

vascular cambium. Subsequently, tissues derived<br />

from the cambium are infected as they are<br />

formed. In large swollen roots, extensive wedgeshaped<br />

masses of hypertrophied medullary ray<br />

tissue may cause the xylem tissue <strong>to</strong> split. At this<br />

stage, the root tissue shows a distinctly mottled<br />

appearance. When the growth of the plasmodia<br />

is complete, they are transformed in<strong>to</strong> masses of<br />

haploid resting spores. Only during the late<br />

stages of resting spore development do the host<br />

nuclei begin <strong>to</strong> degenerate. Eventually, the<br />

resting spores are released in<strong>to</strong> the soil as the<br />

root tissues decay.<br />

3.2.3 Spongospora<br />

The life cycle of S. subterranea, the cause of<br />

powdery scab of pota<strong>to</strong>, is similar <strong>to</strong> that of<br />

P. brassicae (Harrison et al., 1997; Hutchison &<br />

Kawchuk, 1998). Diseased tubers show powdery<br />

pustules at their surface, containing masses of<br />

resting spores clumped in<strong>to</strong> hollow balls. The<br />

resting spores release anisokont zoospores which<br />

can infect the root hairs of pota<strong>to</strong> or <strong>to</strong>ma<strong>to</strong><br />

plants. In the root hairs, plasmodia form which<br />

develop in<strong>to</strong> zoosporangia. Zoospores from such<br />

zoosporangia are capable of infection, resulting<br />

in a further crop of zoosporangia. Zoospores<br />

released from the zoosporangia have also been<br />

observed <strong>to</strong> fuse in pairs or occasionally in<br />

groups of three <strong>to</strong> form quadri- or hexaflagellate<br />

swarmers, but whether these represent true<br />

sexual fusion stages is uncertain. Spongospora<br />

nasturtii causes a disease of watercress in which<br />

the most obvious symp<strong>to</strong>m is a coiling or bending<br />

of the roots. Zoosporangia and resting spore<br />

balls are found in infected root cells (Fig. 3.6),<br />

and plasmodia can migrate through the root<br />

tissue by breaking through host cell walls<br />

(Clax<strong>to</strong>n et al., 1996; Clay & Walsh, 1997). The<br />

encounter of two plasmodia might initiate<br />

sexual reproduction and thus complete the life<br />

cycle without any need for the parasite <strong>to</strong> leave<br />

the host (Heim, 1960).<br />

Fig 3.6 Spongospora nasturtii. Spore balls from watercress<br />

roots with crook root disease.


62 PROTOZOA: PLASMODIOPHOROMYCOTA<br />

In addition <strong>to</strong> being the causal agent of powdery<br />

scab of pota<strong>to</strong>es, S. subterranea is also important<br />

as the vec<strong>to</strong>r of pota<strong>to</strong> mop-<strong>to</strong>p virus disease,<br />

which can reduce the yield of tubers by over 20%<br />

in some varieties (Campbell, 1996; Harrison et al.,<br />

1997). The virus is transmitted by the zoospores<br />

and can also persist for several years in spore balls<br />

in the soil. It seems <strong>to</strong> be located inside the resting<br />

spores (Merz, 1997). Zoospores of S. subterranea<br />

can cause zoosporangial infections in the root<br />

hairs of a wide range of host plants outside<br />

the family Solanaceae, and can transmit viruses<br />

<strong>to</strong> them. Thus S. subterranea and numerous wild<br />

plants can provide a reservoir of infection for<br />

the pota<strong>to</strong> mop-<strong>to</strong>p virus even if pota<strong>to</strong>es have<br />

not been grown in a field for many years. Other<br />

members of the Plasmodiophorales also act as<br />

vec<strong>to</strong>rs for plant viruses, notably Polymyxa betae<br />

which transmits the beet necrotic yellow vein<br />

virus, and P. graminis which transmits several<br />

mosaic viruses on most major cereal crops.<br />

3.3 Control of diseases caused by<br />

Plasmodiophorales<br />

3.3.1 Club root<br />

The control of club root disease is difficult.<br />

Because resting spores retain their viability in<br />

the soil for up <strong>to</strong> 20 years, short-term crop<br />

rotation will not eradicate the disease. The fact<br />

that Plasmodiophora brassicae can infect brassicaceous<br />

weeds such as shepherd’s purse (Capsella<br />

bursa-pas<strong>to</strong>ris) or thalecress (Arabidopsis thaliana)<br />

suggests that the disease can be carried over on<br />

such hosts and that weed control is important.<br />

Moreover, it is known that root hair infection<br />

can also occur on many ubiqui<strong>to</strong>us nonbrassicaceous<br />

hosts such as Papaver and Rumex,<br />

or the grasses Agrostis, Dactylis, Holcus and Lolium.<br />

All infections of non-brassicaceous hosts are<br />

probably reduced <strong>to</strong> the zoosporangial cycle,<br />

and no root clubs are formed. Whether such<br />

infections play any part in maintaining the<br />

disease in the prolonged absence of a brassicaceous<br />

host is not known.<br />

General measures aimed at mitigating the<br />

incidence of clubroot traditionally include<br />

improved drainage and the application of lime,<br />

which retards the primary infection of root hairs.<br />

Since the effect of liming does not persist, it is<br />

possible that it may simply delay the germination<br />

of resting spores and thus prolong their<br />

existence in the soil (Macfarlane, 1952). More<br />

recently, boron added at 10 20 mg kg 1 soil in<br />

conjunction with a high soil pH has been shown<br />

<strong>to</strong> suppress primary as well as secondary infections<br />

(M. A. Webster & Dixon, 1991). Early infection<br />

of seedlings can result in particularly severe<br />

symp<strong>to</strong>ms, so it is important <strong>to</strong> raise seedlings in<br />

non-infected or steam-sterilized soil. The young<br />

plants can then be transplanted <strong>to</strong> infested<br />

soil. Since it is known that some resting spores<br />

survive animal digestion, manure from animals<br />

fed with diseased material should not be used<br />

for growing brassicas.<br />

Infection can be retarded by the application<br />

of mercury-containing compounds or benomyl,<br />

but these are now banned in many countries. At<br />

present, no economically and ecologically acceptable<br />

fungicide appears <strong>to</strong> be available, although<br />

research efforts continue (Mitani et al., 2003).<br />

Some attempts have been made <strong>to</strong> establish biological<br />

control methods for P. brassicae (Narisawa<br />

et al., 1998; Tils<strong>to</strong>n et al., 2002), but it is doubtful<br />

whether such methods will gain full commercial<br />

viability in the near future.<br />

In recent years, increasing emphasis has been<br />

placed on breeding club root resistant cultivars<br />

of crop plants. The weed Arabidopsis thaliana,<br />

which develops the full set of club root symp<strong>to</strong>ms,<br />

has been used as a host for such studies<br />

because it is accessible by molecular biological<br />

methods. Natural resistance in Arabidopsis is<br />

based on a single gene and involves the hypersensitive<br />

response, in which infected plant<br />

cells die before the pathogen has had a chance<br />

<strong>to</strong> multiply. The resistance of susceptible cultivars<br />

can be enhanced by transformation with<br />

various resistance genes, e.g. a gene from<br />

mistle<strong>to</strong>e (Viscum album) encoding visco<strong>to</strong>xin,<br />

a thionin-type cystein-rich polypeptide with<br />

antimicrobial activity (Hol<strong>to</strong>rf et al., 1998).<br />

Further, mutant lines with reduced levels of<br />

IAA precursors show reduced club development<br />

(Ludwig-Müller, 1999).


CONTROL OF DISEASES CAUSED BY PLASMODIOPHORALES<br />

63<br />

In contrast <strong>to</strong> Arabidopsis, natural resistance<br />

in cabbage is multigenic, with no obvious<br />

hypersensitive response (Ludwig-Müller, 1999).<br />

Breeding for resistance is difficult (Bradshaw<br />

et al., 1997) and may not provide long-lasting<br />

success due <strong>to</strong> the development of new virulent<br />

races of P. brassicae on the resistant cultivars after<br />

a few years in the field. By 1975, 34 different<br />

physiological races of P. brassicae from Europe<br />

had already been differentiated based on infection<br />

experiments with Brassica cultivars varying<br />

in their degree of resistance (Buczacki et al.,<br />

1975). Further, P. brassicae can still infect root<br />

hairs and reproduce by zoosporangia even in<br />

resistant cultivars.<br />

3.3.2 Powdery scab and crook root<br />

Powdery scab of pota<strong>to</strong>es is normally of relatively<br />

slight economic importance and amelioration of<br />

the disease can be brought about by good drainage.<br />

Pota<strong>to</strong> mop-<strong>to</strong>p virus infections can be more<br />

serious, however. Transgenic plants containing<br />

the viral coat protein gene have been shown <strong>to</strong> be<br />

completely resistant against infections by the<br />

virus (Reavy et al., 1995), and it may be possible <strong>to</strong><br />

produce transgenic crop plants in future.<br />

Crook root of watercress can be controlled by<br />

application of zinc <strong>to</strong> the water supply. The zinc<br />

can be applied by dripping zinc sulphate in<strong>to</strong><br />

the irrigation water for watercress beds <strong>to</strong> give<br />

a final concentration of about 0.5 ppm, or by the<br />

Fig 3.7 Hap<strong>to</strong>glossa heteromorpha parasitizing nema<strong>to</strong>des. (a) Single young thallus in a dead nema<strong>to</strong>de. (b) Single maturing<br />

sporangium with developing dome-shaped exit papillae. (c) Nema<strong>to</strong>de body containing several plasmodia and sporangia.<br />

One sporangium has released large aplanospores, and an adjacent one small ones. (d) Small aplanospores, one germinating <strong>to</strong><br />

form a gun cell. (e) Large aplanospores, one germinating <strong>to</strong> form a gun cell. (a c) <strong>to</strong> same scale; d,e <strong>to</strong> same scale. Redrawn from<br />

Glockling and Beakes (2000a).


64 PROTOZOA: PLASMODIOPHOROMYCOTA<br />

addition of finely powdered glass containing<br />

zinc oxide (zinc frit) <strong>to</strong> the beds. The slow<br />

release of zinc from the frits maintains a sufficiently<br />

high concentration <strong>to</strong> inhibit infection<br />

(Tomlinson, 1958).<br />

3.4 Hap<strong>to</strong>glossa<br />

(Hap<strong>to</strong>glossales)<br />

3.4.1 General biological features of<br />

Hap<strong>to</strong>glossa<br />

If a slurry of soil or herbivore dung is spread on a<br />

weak medium such as tap water agar or cornmeal<br />

agar, the nema<strong>to</strong>des or rotifers contained within<br />

these samples may become parasitized and killed<br />

by fungi producing thalli within the cadavers.<br />

Although superficially resembling the plasmodia<br />

of Plasmodiophora, this term cannot be applied <strong>to</strong><br />

Hap<strong>to</strong>glossa because its thalli are surrounded by a<br />

wall at all stages of development. One or several<br />

thalli may fill almost the entire body cavity of<br />

a nema<strong>to</strong>de and become converted in<strong>to</strong> sporangia<br />

upon maturity (Fig. 3.7). Sporangia of some<br />

species of Hap<strong>to</strong>glossa release zoospores which<br />

are anisokont, with both flagella of the whiplash<br />

type. Zoospore release occurs through one or<br />

several exit papillae (Barron, 1977). Zoospores of<br />

Hap<strong>to</strong>glossa are weak swimmers and encyst within<br />

a few minutes in the vicinity of the host cadaver<br />

from which they were released. Other species<br />

of Hap<strong>to</strong>glossa do not release zoospores but<br />

produce non-motile spores (aplanospores) resembling<br />

cysts of the zoospore-forming species.<br />

Aplanospore release occurs by explosive rupture<br />

of the exit tube, followed by several further,<br />

progressively weaker bursts of discharge<br />

(Glockling & Beakes, 2000a). A few hours after<br />

their formation or release, cysts or aplanospores<br />

germinate <strong>to</strong> produce an elongated or glossoid<br />

(¼ <strong>to</strong>ngue-shaped) cell, which is also often called<br />

a gun cell or an infection cell. This explosively<br />

injects a small amount of walled pro<strong>to</strong>plasm<br />

(sporidium) containing a nucleus and a few<br />

organelles in<strong>to</strong> a host passing by (see below). The<br />

sporidium enlarges <strong>to</strong> form a new thallus and,<br />

upon host death, a new sporangium. The mechanism<br />

of gun cell discharge is rather similar <strong>to</strong><br />

that found in cysts of Plasmodiophora or Polymyxa.<br />

This, <strong>to</strong>gether with the occurrence of anisokont<br />

zoospores, has been taken as an indication that<br />

Hap<strong>to</strong>glossa should be included in the Plasmodiophoromycota<br />

(Beakes & Glockling, 1998; Dick,<br />

2001a), whereas formerly the genus was thought<br />

<strong>to</strong> be related <strong>to</strong> the Oomycota.<br />

The aplanosporic species of Hap<strong>to</strong>glossa produce<br />

spores of two distinctly different sizes,<br />

Fig 3.8 Hap<strong>to</strong>glossa sp. (a) Tip of a developing gun cell.<br />

The muzzle is still sealed by its plug (Pl). Bore (Bo) and<br />

needle chamber (NC) are visible. (b) Transmission<br />

electron micrograph of a mature gun cell.The basal part<br />

of the gun cell is entirely occupied by the enlarging<br />

posterior vacuole (Vac).Original prints kindly supplied by<br />

S.L.Glockling.


HAPTOGLOSSA (HAPTOGLOSSALES)<br />

65<br />

although any one sporangium produces propagules<br />

only of either size (Glockling & Beakes,<br />

2000a; Fig. 3.7). In contrast <strong>to</strong> the Plasmodiophorales,<br />

sexual reproduction or resting stages<br />

have not yet been described for any species<br />

of Hap<strong>to</strong>glossa, and it is difficult at present <strong>to</strong><br />

explain the occurrence of spores of different<br />

sizes. What appears clear is that each thallus is<br />

the result of a discrete infection event.<br />

3.4.2 The gun cell of Hap<strong>to</strong>glossa<br />

Germination of the spherical zoospore cyst or<br />

aplanospore of Hap<strong>to</strong>glossa occurs by means of<br />

a short germ tube which enlarges <strong>to</strong> form the<br />

elongated gun cell (Robb & Lee, 1986a). This<br />

remains attached <strong>to</strong> the cyst until maturity and<br />

is perched on <strong>to</strong>p of it in many species. The<br />

mature gun cell (Figs. 3.8, 3.9a) shows strong<br />

ultrastructural similarities <strong>to</strong> the infection<br />

apparatus of Plasmodiophora (see Fig. 3.5) and is<br />

the object of considerable mycological curiosity.<br />

A tube leads in<strong>to</strong> the pointed tip of the gun cell<br />

but its opening (muzzle) is separated from the<br />

exterior by a thin wall (plug) for most of its<br />

development (Fig. 3.8a). The formation of this<br />

internal tube from the tip of the gun cell<br />

backwards has been likened <strong>to</strong> inverted internal<br />

tip growth and is mediated by a scaffold of actin<br />

fibres against the turgor pressure of the gun cell<br />

(Beakes & Glockling, 1998). The inner (noncy<strong>to</strong>plasmic)<br />

surface of the anterior part of<br />

the tube (bore) is lined with fibrillar material.<br />

A second wall separates the bore from a swollen<br />

section of the tube, the needle chamber. This<br />

contains a projectile (needle) resembling the<br />

bullet of Plasmodiophora, but terminating in a<br />

much finer tip, possibly reflecting the different<br />

properties of the host surface which it has <strong>to</strong><br />

puncture. The needle is held in place by a<br />

complex set of cones and cylinders (Fig. 3.8a)<br />

which are thought <strong>to</strong> exercise a restraining<br />

function, fixing the needle against the high<br />

turgor pressure of the gun cell. The cones and<br />

cylinders may contain actin filaments. The shaft<br />

of the needle is much wider than its tip. The<br />

posterior (innermost) part of the tube (tail) coils<br />

around itself and the nucleus, almost <strong>to</strong>uching<br />

the side of the needle chamber. The tail is walled,<br />

Fig 3.9 Schematic drawings of the nema<strong>to</strong>de penetration<br />

mechanism in Hap<strong>to</strong>glossa.(a)Guncellreadyfordischarge.The<br />

tube has already protruded <strong>to</strong> form a beak (Bk), the exterior of<br />

which is lined by a glue originating from the inside surface of<br />

the bore.This aids in the attachment of the gun cell <strong>to</strong> a passing<br />

nema<strong>to</strong>de.The needle (Ne) is held in position by actin<br />

filaments inside the needle chamber (NC), which is separated<br />

from the outside by a wall. Behind the needle chamber is the<br />

coiled tail (T) which contains wallmaterialin its lumen (dotted<br />

area). In fact, the tail is multi-layered, but this has not been<br />

illustrated here.The tail coils round the nucleus (Nuc) and a<br />

Golgi stack (G), and mi<strong>to</strong>chondria (Mit) are also located in the<br />

vicinity.The posterior of the gun cell is filled by one large<br />

vacuole (Vac). (b) Tip of a fired gun cell showing the everted<br />

tail which has penetrated the nema<strong>to</strong>de cuticle and has<br />

formed a sporidium inside the nema<strong>to</strong>de body (above the<br />

cuticle).The wall material formerly located inside the tail has<br />

formed the sporidium wall.The detached needle is also visible<br />

inside the nema<strong>to</strong>de body.For a more detailed description of<br />

the eversion process, see Glockling and Beakes (2000b).


66 PROTOZOA: PLASMODIOPHOROMYCOTA<br />

and additional electron-dense cell wall precursor<br />

material is deposited within the lumen of the<br />

tail. Synthesis of the tube is mediated by one<br />

large Golgi stack which is always closely associated<br />

with the nucleus and faces the inwardgrowing<br />

tube tip, emitting vesicles <strong>to</strong>wards it. As<br />

the tube extends and coils round the nucleus,<br />

the nucleus and Golgi stack turn like a dial by<br />

360° (Beakes & Glockling, 1998, 2000). The turgor<br />

pressure of the gun cell is probably generated by<br />

a large posterior vacuole (Fig. 3.8b), similar <strong>to</strong><br />

that found in cysts of Plasmodiophora. The<br />

osmotically active solutes required for turgor<br />

generation may originate from the degradation<br />

of lipid droplets within the enlarging vacuole.<br />

Shortly before discharge, the increasing<br />

turgor pressure of the posterior vacuole is<br />

thought <strong>to</strong> push the tip of the gun cell forward;<br />

the wall sealing the muzzle is lost, and the bore<br />

shortens and extends a beak-like projection<br />

(Fig. 3.9a). The cell wall material from the<br />

interior of the bore now forms the external<br />

beak wall, and the needle is ready for injection.<br />

The nature of the discharge trigger probably<br />

varies between different species of Hap<strong>to</strong>glossa<br />

and may be chemical or mechanical. The beak<br />

wall is thought <strong>to</strong> act as an adhesive and<br />

immediately glues the gun cell <strong>to</strong> the cuticle of<br />

a passing nema<strong>to</strong>de or rotifer. Firm attachment<br />

is necessary <strong>to</strong> provide resistance against the<br />

recoil of the needle attempting <strong>to</strong> penetrate<br />

the <strong>to</strong>ugh cuticle of the host, as it is for the<br />

penetrating bullet in adhesoria of Plasmodiophora.<br />

Beakes and Glockling (1998) speculated<br />

that stretch-activated membrane channels<br />

(see p. 8) might be involved in triggering the<br />

launch of the needle. Following attachment,<br />

Ca 2þ ions entering the needle chamber would<br />

cause the actin-rich cones and cylinders near the<br />

needle tip <strong>to</strong> contract and rupture. Once the<br />

constraints exercised by the cones and cylinders<br />

are broken, the high turgor pressure of the gun<br />

cell will immediately fire the needle, followed by<br />

explosive eversion of the entire tube which forms<br />

a syringe, conducting the nucleus, Golgi apparatus<br />

and mi<strong>to</strong>chondria of the gun cell through<br />

the nema<strong>to</strong>de cuticle (Fig. 3.9b). The infective<br />

propagule is called a sporidium because it is<br />

surrounded by a wall, the material for which is<br />

probably contributed by precursor material at<br />

the end of the tail section (Robb & Lee, 1986b;<br />

Glockling & Beakes, 2000b).


4<br />

Straminipila: minor fungal phyla<br />

4.1 <strong>Introduction</strong><br />

The kingdom Chromista was erected by Cavalier-<br />

Smith (1981, 1986) <strong>to</strong> accommodate eukaryotic<br />

organisms which are distinguishable from the<br />

Pro<strong>to</strong>zoa by a combination of characters. Some<br />

of these are concerned with details of pho<strong>to</strong>synthesis,<br />

such as the enclosure of chloroplasts<br />

in sheets of endoplasmic reticulum, and the<br />

absence of chlorophyll b, the latter feature<br />

being used for the naming of the kingdom.<br />

Other defining characters apply also <strong>to</strong> the nonpho<strong>to</strong>synthetic<br />

members of the Chromista (Kirk<br />

et al., 2001). These are as follows:<br />

1. The structural cell wall polymer is cellulose,<br />

in contrast <strong>to</strong> walls of Eumycota which<br />

contain chitin.<br />

2. The inner mi<strong>to</strong>chondrial membrane is<br />

folded in<strong>to</strong> tubular cristae (Fig. 4.1a) which are<br />

also found in plants. In contrast, mi<strong>to</strong>chondrial<br />

cristae are generally lamellate in the kingdoms<br />

Eumycota (Fig. 4.1b) and Animalia.<br />

3. Golgi stacks (dictyosomes) are present;<br />

these are also found in the Pro<strong>to</strong>zoa (see p. 64).<br />

In contrast, in the Eumycota the Golgi apparatus<br />

is usually reduced <strong>to</strong> single cisternae (see<br />

Figs. 1.3, 1.10).<br />

4. Flagella are usually present during particular<br />

stages of the life cycle; they always include<br />

one straminipilous flagellum (Lat. stramen ¼<br />

straw, pilus ¼ hair). Dick (2001a) considered<br />

this feature <strong>to</strong> be of such high phylogenetic<br />

significance that he has renamed the kingdom<br />

Chromista as Straminipila. The straminipilous<br />

flagellum is discussed in detail in the following<br />

section.<br />

5. The amino acid lysine is synthesized via<br />

the a,e-diaminopimelic acid (DAP) pathway.<br />

Diaminopimelic acid originates from aspartic<br />

semialdehyde and pyruvic acid and is present<br />

in terrestrial plants, green algae, Chromista<br />

and prokaryotes. The alternative route, the<br />

a-aminoadipic acid (AAA) pathway, draws on<br />

a-ke<strong>to</strong>glutaric acid and acetyl-CoA and is found<br />

almost exclusively in members of the Eumycota.<br />

Yet other organisms, including animals and<br />

Pro<strong>to</strong>zoa, are auxotrophic for lysine (Griffin,<br />

1994). Lysine biosynthesis has been used as a<br />

chemotaxonomic marker for some time (Vogel,<br />

1964; LéJohn, 1972).<br />

The kingdom Chromista/Straminipila currently<br />

includes the dia<strong>to</strong>ms, golden and<br />

brown algae, chrysophytes and cryp<strong>to</strong>monads,<br />

as well as three phyla of straminipilous organisms<br />

traditionally studied by mycologists,<br />

i.e. the Oomycota, Hyphochytriomycota and<br />

Labyrinthulomycota. The first two groups are<br />

also called straminipilous fungi because of the<br />

similarity of their mode of life <strong>to</strong> the fungal<br />

lifestyle (Dick, 2001a). The Oomycota are by far<br />

the more important of these, and are considered<br />

in detail in Chapter 5. The Hyphochytriomycota<br />

and Labyrinthulomycota are treated briefly in<br />

the present chapter. The Straminipila as circumscribed<br />

above are a diverse but natural


68 STRAMINIPILA: MINOR FUNGAL PHYLA<br />

Fig 4.1 Mi<strong>to</strong>chondrial ultrastructure observed by<br />

transmission electron microscopy. (a) Mi<strong>to</strong>chondrion of<br />

Phy<strong>to</strong>phthora erythroseptica (Oomycota).The inner<br />

mi<strong>to</strong>chondrial membrane is folded in<strong>to</strong> a complex tubular<br />

network. (b) Mi<strong>to</strong>chondrion of Sordaria fimicola (Ascomycota)<br />

with the inner membrane appearing lamellate. Mi<strong>to</strong>chondrial<br />

ribosomes (arrows) are also visible. Reprinted from Weber<br />

et al. (1998), with permission from Elsevier.<br />

(monophyletic) grouping which has been<br />

confirmed by comparisons of the small-subunit<br />

(18S) ribosomal DNA sequences (e.g. Hausner<br />

et al., 2000; Fig. 4.2).<br />

4.2 The straminipilous<br />

flagellum<br />

The eukaryotic flagellum is a highly conserved<br />

structure. It is formed within the cy<strong>to</strong>plasm by a<br />

kine<strong>to</strong>some, i.e. a microtubule-organizing centre<br />

resembling the centriole which co-ordinates the<br />

formation of the microtubular spindle during<br />

nuclear division. Like the centriole, the kine<strong>to</strong>some<br />

contains an outer ring of nine triplets<br />

of microtubules surrounding two central microtubules<br />

(see Figs. 6.2 and 6.19). The flagellum<br />

extends outwards from the centriole as nine<br />

doublets of microtubules surrounding the two<br />

single central microtubules. This is the 9 þ 2<br />

arrangement. Where the eukaryotic flagellum<br />

protrudes beyond the cell surface, it is<br />

ensheathed by the plasma membrane. Within<br />

the flagellum, there are no obvious cy<strong>to</strong>plasmic<br />

features other than the microtubules which<br />

<strong>to</strong>gether are called the axoneme. Flagella<br />

which are entirely smooth or bear a coat of<br />

fine fibrillar surface material visible only by<br />

high-resolution electron microscopy (Fig. 4.3a;<br />

Andersen et al., 1991) are commonly called<br />

whiplash flagella. Dick (2001a) has pointed out<br />

that whiplash flagella in a strict sense are<br />

pointed at their tip due <strong>to</strong> the fact that the two<br />

inner microtubules are longer than the nine<br />

outer doublets (Fig. 4.3a).<br />

A second type of flagellum is decorated with<br />

hair-like structures 1 2 mm long (Fig. 4.3b). This<br />

is the tinsel or straminipilous flagellum (Dick,<br />

1997). The hairs are called tripartite tubular<br />

hairs (TTHs) because they are divided in<strong>to</strong> three<br />

parts. They were formerly called mastigonemes,<br />

thereby naming the fungi which produced them<br />

Mastigomycotina, but both terms are no longer<br />

used. Each TTH is attached <strong>to</strong> the flagellum by a<br />

conical base pointed <strong>to</strong>wards the axoneme.<br />

The main part of the TTH is a long tubular<br />

shaft thought <strong>to</strong> consist of two fibres of different<br />

thickness coiled around each other (Domnas<br />

et al., 1986). At the tip of the TTH, the two<br />

fibres separate from each other <strong>to</strong> form loose<br />

ends (Figs. 4.3b, 4.4). In the TTHs of some<br />

straminipilous organisms, only one loose end is<br />

visible (Fig. 4.7b). TTHs are assembled in antiparallel<br />

arrays in Golgi-derived vesicles of the<br />

maturing zoospore, and are released by fusion of<br />

the vesicles with the plasma membrane (Fig. 4.5;<br />

Heath et al., 1970; Cooney et al., 1985). When a<br />

spore encysts, the flagellum may be withdrawn,<br />

shed or coiled around the spore. If it is withdrawn,<br />

the TTHs are sloughed off and left behind<br />

as a tuft on the surface of the cyst (Dick, 1990a).<br />

TTHs are arranged in two rows along the<br />

axoneme. The cones of each row are adjacent <strong>to</strong><br />

an outer microtubule doublet, and because there<br />

are nine such doublets, the two rows of TTHs<br />

are at an angle of about 160° rather than 180°<br />

<strong>to</strong> each other (Fig. 4.4a). In zoospores of


THE STRAMINIPILOUS FLAGELLUM<br />

69<br />

Fig 4.2 Unrooted phylogenetic tree of the Straminipila and members of other kingdoms, based on analyses of18S rDNA<br />

sequences.Redrawn and modified from Hausner et al. (2000), by copyright permission of the National Research Council of Canada.<br />

Fig 4.3 Ultrastructure of flagella in Straminipila. (a) Whiplash flagellum of Pythium monospermum (Oomycota).The tip is narrower<br />

than the main body of the flagellum because the two central microtubules are longer than the nine outer doublets. Arrows indicate<br />

the coating of the flagellum with very fine hairs. (b) Tinsel flagellum of Achlya colorata (Oomycota) with numerousTTHs. EachTTH<br />

ends in two fibres, one longer and thicker than the other (arrows).Original images kindly provided by M.W. Dick and I.C. Hallett.


70 STRAMINIPILA: MINOR FUNGAL PHYLA<br />

Fig 4.4 Organization of the straminipilous flagellum.<br />

(a) Postulated attachment of TTHs <strong>to</strong> the microtubule<br />

doublets1and 5 of the axoneme as seen in transverse section<br />

(after Dick, 2001a). (b) Longitudinal arrangement of TTHs<br />

along the axoneme of a straminipilous flagellum.Only one row<br />

of TTHs is drawn.TheTTHs are thought <strong>to</strong> be arranged in an<br />

alternating fashion as regards the orientation of long and<br />

short fibres in adjacent TTHs. b redrawn from Dick (1990a).<br />

ß 1990 Jones and Bartlett Publishers, Sudbury, MA.<br />

www.jbpub.com.<br />

straminipilous fungi, the straminipilous flagellum<br />

always seems <strong>to</strong> point <strong>to</strong>wards the direction<br />

of movement, and Dick (1990a, 2001a) has<br />

advanced a theory <strong>to</strong> explain how movement<br />

can be generated from a sinusoidal wave starting<br />

at the flagellar base, likening the straminipilous<br />

flagellum <strong>to</strong> ‘a rowing eight with fixed oars and<br />

a flexible keel’ (Fig. 4.4b; Dick, 2001a). An anterior<br />

straminipilous flagellum therefore pulls the<br />

spore through the water, whereas a backwardly<br />

directed whiplash flagellum pushes the spore.<br />

The construction of the straminipilous flagellum<br />

is so elaborate that it is most unlikely <strong>to</strong><br />

have arisen more than once during evolution<br />

(Dick, 2001a). The presence of a straminipilous<br />

flagellum, whether or not accompanied by<br />

another, smooth flagellum, therefore indicates<br />

membership in the Straminipila.<br />

4.3 Hyphochytriomycota<br />

This group, formerly called Hyphochytridiomycetes<br />

probably due <strong>to</strong> the perpetuation of<br />

Fig 4.5 Schematic drawing of a L.S. of a zoospore of the<br />

hyphochytrid Hyphochytrium catenoides.The elongated shape<br />

of the zoospore and of the nucleus (N) is maintained by a<br />

system of ‘rootlets’consisting of parallel bundles of<br />

micro<strong>to</strong>bules (thick lines).The straminipilous flagellum arises<br />

from a kine<strong>to</strong>some (Kin). A second, non-functional<br />

kine<strong>to</strong>some (NFK) is interpreted as the base of a whiplash<br />

flagellum lost in the course of evolution from a heterokont<br />

ances<strong>to</strong>r. Mi<strong>to</strong>chondria (Mit),TTH-containing vesicles (TV),<br />

a Golgi stack (G), ER, ribosomes (Rib), a large basal lipid<br />

droplet (LD) and microbodies (MB) are also visible. Some<br />

organelles of unknown function, e.g. electron-opaque bodies<br />

and osmiophilic bodies, have been omitted from the<br />

original for improved clarity. Redrawn and modified from<br />

Cooney et al. (1985).<br />

a typographical error (see Dick, 1983), is a very<br />

small phylum currently comprising 23 species in<br />

6 genera (Kirk et al., 2001). The Hyphochytriomycota<br />

(colloquially called hyphochytrids) are


LABYRINTHULOMYCOTA<br />

71<br />

phylogenetically closely related <strong>to</strong> the Oomycota<br />

(van der Auwera et al., 1995; Hausner et al., 2000;<br />

see Fig. 4.2). Treatments of the group have been<br />

given by Karling (1977), Fuller (1990, 2001) and<br />

Dick (2001a). The diagnostic feature is the<br />

zoospore with its single anterior straminipilous<br />

flagellum (Fig. 4.5). This kind of zoospore is not<br />

found in any other known life form. The<br />

zoospore of hyphochytrids contains one prominent<br />

Golgi stack, one nucleus, and lipid droplets<br />

and microbodies (Barr & Allan, 1985; Cooney<br />

et al., 1985). The latter are not arranged in a<br />

microbody lipid complex like they are in<br />

chytrids (cf. Fig. 6.3). The TTHs are localized<br />

within Golgi-derived vesicles. The flagellum<br />

arises from a kine<strong>to</strong>some, with microtubules<br />

rooting deeply within the spore and probably<br />

maintaining its shape. A second (dormant)<br />

kine<strong>to</strong>some lies adjacent but at an angle, at the<br />

same position as that which gives rise <strong>to</strong> the<br />

backward-directed smooth flagellum in zoospores<br />

of Oomycota. This whiplash flagellum is<br />

missing in Hyphochytriomycota, and Barr and<br />

Allan (1985) have speculated that it could have<br />

been lost during evolution of the latter from the<br />

former. Like the Oomycota, hyphochytrids<br />

synthesize lysine by the a,e-diaminopimelic acid<br />

(DAP) pathway (Vogel, 1964).<br />

Hyphochytrids occur in the soil and in<br />

aquatic environments (both freshwater and<br />

marine) as saprotrophs or parasites of algae,<br />

oospores of Oomycota or azygospores of<br />

Glomales. Hyphochytrium peniliae was reported<br />

once as the cause of a devastating epidemic of<br />

marine crayfish (Artemchuk & Zelezinkaya,<br />

1969), but no further cases have been observed<br />

since. Some species can be isolated in<strong>to</strong> pure<br />

culture relatively easily (Fuller, 1990).<br />

Zoospores encyst by withdrawing their flagellum<br />

and secreting a wall, leaving the TTHs<br />

dispersed on the surface of the cyst wall<br />

(Beakes, 1987). The cyst germinates by enlargement<br />

or by putting out rhizoids. Because of the<br />

similarity of their vegetative thalli with those of<br />

Chytridiomycota (see Chapter 6), hyphochytrids<br />

have been studied primarily by comparison with<br />

chytrids, and the same terminology has been<br />

used (see Fig. 6.1). Depending on the species,<br />

cysts germinate <strong>to</strong> develop in three different<br />

ways, which have been used <strong>to</strong> subdivide<br />

the Hyphochytriomycota in<strong>to</strong> families:<br />

(1) Holocarpic thalli are produced by simple<br />

enlargement of the cyst. The entire content of<br />

the sac-like thallus ultimately becomes converted<br />

in<strong>to</strong> zoospores (Anisolpidiae, e.g. Anisolpidium<br />

which parasitizes marine algae; Canter, 1950).<br />

(2) In eucarpic monocentric thalli, the cyst<br />

produces a bunch of rhizoids at one end, which<br />

anchor the enlarging thallus <strong>to</strong> the substratum<br />

and/or absorb nutrients (Rhizidiomycetidae, e.g.<br />

Rhizidiomyces; Wynn & Ep<strong>to</strong>n, 1979). (3) In<br />

eucarpic polycentric thalli, a broad hypha-like<br />

germ tube emerges, branches and produces<br />

several zoosporangia (Hyphochytriaceae, e.g.<br />

Hyphochytrium; Ayers & Lumsden, 1977). The<br />

asexual life cycle is completed when a fresh<br />

crop of zoospores is released. Sexual reproduction<br />

has not yet been reliably described for the<br />

hyphochytrids.<br />

4.4 Labyrinthulomycota<br />

Whereas the Hyphochytriomycota described in<br />

the previous section have a strong resemblance<br />

<strong>to</strong> true fungi (especially Chytridiomycota),<br />

the Labyrinthulomycota do not, and the<br />

only justification for mentioning them here is<br />

the fact that they have traditionally been<br />

studied by mycologists. They have been the<br />

subject of numerous taxonomic rearrangements,<br />

and are known under many different names such<br />

as Labyrinthomorpha, Labyrinthista and<br />

Labyrinthulea. Some 48 species are currently<br />

recognized (Kirk et al., 2001). DNA sequence<br />

comparisons have placed them within the<br />

Straminipila (Fig. 4.2; Hausner et al., 2000;<br />

Leander & Porter, 2001), and they are characterized<br />

by having heterokont flagellation, i.e.<br />

possessing a straminipilous and a whiplash<br />

flagellum with a pointed tip (Fig. 4.7). In addition,<br />

they have mi<strong>to</strong>chondria with tubular cristae.<br />

Recent treatments of this group can be found in<br />

Moss (1986), Porter (1990) and Dick (2001a).<br />

Labyrinthulomycota occur in freshwater and<br />

marine environments where they are attached <strong>to</strong><br />

solid substrata by means of networks of slime


72 STRAMINIPILA: MINOR FUNGAL PHYLA<br />

within which individual vegetative cells are<br />

contained. For this reason, they are sometimes<br />

referred <strong>to</strong> as ‘slime nets’ (Porter, 1990). The<br />

vegetative cells possess a wall which, uniquely,<br />

is produced from Golgi-derived scales of a<br />

polymer of l-galac<strong>to</strong>se (Dick, 2001a). These<br />

scales are located between the plasma membrane<br />

and the inner membrane of the slime net. The<br />

slime net is delimited by an inner and an outer<br />

membrane and is produced by specialized organelles<br />

termed sagenogens or bothrosomes; the<br />

net membranes are continuous with the plasma<br />

membrane at the sagenogen (Perkins, 1972).<br />

Labyrinthulomycota feed by absorption (osmotrophy)<br />

of nutrients. The nets contain degradative<br />

enzymes which can lyse plant material or<br />

microbial cells. Two orders are distinguished.<br />

4.4.1 Labyrinthulales<br />

Members of this order, especially of the genus<br />

Labyrinthula, can be readily isolated from marine<br />

angiosperms such as Zostera and Spartina or from<br />

seaweed by placing a small piece of one of these<br />

substrata directly on low-nutrient sea water agar<br />

augmented with penicillin and strep<strong>to</strong>mycin<br />

(Porter, 1990). Within a few days, a fine network<br />

of strands can be seen extending over the agar<br />

surface (Fig. 4.6). Labyrinthula spp. can be kept in<br />

monoxenic culture with yeasts or bacteria as<br />

food source. These are presumably lysed by the<br />

enzymes contained in the slime net.<br />

A closer examination shows that the network<br />

consists of branched slime tubes within which<br />

spindle-shaped cells move backwards and<br />

forwards (Fig. 4.7a; see Webster, 2006a).<br />

Movement of a speed up <strong>to</strong> 100 mm min 1 has<br />

been reported and is due <strong>to</strong> a system of<br />

contractile actin-like proteins in the slime net<br />

(Nakatsuji & Bell, 1980). Cells occasionally aggregate<br />

<strong>to</strong> form sporangia containing numerous<br />

round cysts. Following meiosis, eight heterokont<br />

zoospores (Figs. 4.6a, 4.7b) are released by each<br />

cyst. These possess a pigmented eyespot not<br />

found in other types of heterokont zoospore<br />

(Porter, 1990). It is, however, unclear whether<br />

zoospores can establish new colonies (Porter,<br />

1990). Asexual reproduction occurs by division<br />

of spindle cells within the slime net, and<br />

fragments of such a colony can establish new<br />

colonies (Porter, 1972). Further details of the life<br />

cycle appear <strong>to</strong> be unknown at present.<br />

Labyrinthula spp. were implicated as pathogens<br />

in a wasting epidemic of eelgrass (Zostera<br />

marina) at the west coast of North America in the<br />

1930s (Young, 1943; Muehlstein et al., 1991),<br />

causing considerable disturbance <strong>to</strong> the lit<strong>to</strong>ral<br />

ecosystem and collateral damage <strong>to</strong> the<br />

local fisheries industry. However, although<br />

Labyrinthula spp. are still frequently associated<br />

with pieces of moribund Zostera shoots, no<br />

further epidemics seem <strong>to</strong> have occurred since.<br />

Instead, a new species, L. terrestris, has recently<br />

been identified as the cause of a rapid blight of<br />

Fig 4.6 Labyrinthula. (a) Zoospore with<br />

long anterior straminipilous flagellum<br />

and a short posterior whiplash flagellum<br />

with a pointed tip (after Amon &<br />

Perkins,1968). (b d) Portions of colonies<br />

at different magnifications. In (c) spindle<br />

cells are seen in swellings in the slime<br />

tracks.


LABYRINTHULOMYCOTA<br />

73<br />

Fig 4.7 Ultrastructural features of Labyrinthulomycota. (a) Spindle-shaped cells of Labyrinthula within their slime net. Each cell has<br />

mi<strong>to</strong>chondria with tubular cristae (Mit),Golgi stacks (G), a single nucleus (N), and cortical lipid droplets (LD).The slime net is<br />

produced by several sagenogens (Sag) in each cell.The plasma membrane is continuous with the inner membrane of the slime net.<br />

Wall scales are released at the sagenogen point and accumulate between the plasma membrane and the inner membrane of the<br />

slime net. (b) Biflagellate heterokont zoospore of Labyrinthula showing an eyespot (E) close <strong>to</strong> the base of the whiplash flagellum.<br />

Note that eachTTH of the Labyrinthula zoospore produces only one terminal fibre. (c) Young thallus of Thraus<strong>to</strong>chytrium.<br />

Mi<strong>to</strong>chondria with tubular cristae, a Golgi stack, lipid droplets and larger vacuoles (Vac) are seen.The wall consists of scales<br />

pre-formed in Golgi-derived vesicles (Ves).The slime net is produced at the base of the thallus by a single sagenogen. All images<br />

schematic and not <strong>to</strong> scale; redrawn and modified from Porter (1990). ß1990 Jones and Bartlett Publishers, Sudbury, MA.<br />

www.jbpub.com.<br />

turf-grass on golf courses, infection presumably<br />

being brought about by irrigation with contaminated<br />

water of unusually high salinity (Bigelow<br />

et al., 2005).<br />

4.4.2 Thraus<strong>to</strong>chytriales<br />

Thraus<strong>to</strong>chytrids are probably ubiqui<strong>to</strong>us in<br />

marine environments, occurring on organic<br />

debris as well as calcareous shells of invertebrates<br />

(Porter & Lingle, 1992). Like the labyrinthulids,<br />

they feed on organic matter, algae and<br />

bacteria (Raghukumar, 2002). Thraus<strong>to</strong>chytrids<br />

can be baited by sprinkling pine pollen grains<br />

on<strong>to</strong> water samples or organic debris immersed<br />

in water. Within one <strong>to</strong> several days, the pollen<br />

grains become colonized by one or several thalli,<br />

the main bodies of which protrude beyond the<br />

grain surface (Figs. 4.8a,b). If colonized pollen<br />

grains are transferred <strong>to</strong> a suitable agar medium<br />

containing sea salts, yeast extract and sugar<br />

(Yokochi et al., 1998), thalli will grow on the agar<br />

surface and may be induced <strong>to</strong> release zoospores<br />

by mounting them in water. Thraus<strong>to</strong>chytrids<br />

can be s<strong>to</strong>red in pollen grain suspensions or on<br />

agar overlaid with sea water. They also possess<br />

the ability <strong>to</strong> survive in a dry state at room<br />

temperature for a year or longer (Porter, 1990).<br />

The thallus of thraus<strong>to</strong>chytrids superficially<br />

resembles that of an epibiotic monocentric<br />

chytrid in having a roughly spherical shape<br />

with ‘rhizoids’ at its base (Fig. 4.8c). These<br />

‘rhizoids’ are, in fact, the slime net produced<br />

by one basal sagenogen (Fig. 4.7c). The thallus is<br />

surrounded by Golgi-derived scales forming<br />

a wall, but the slime net does not extend over<br />

the thallus. Sexual reproduction is unknown,<br />

but asexual biflagellate heterokont zoospores<br />

are released from the main body of the thallus,


74 STRAMINIPILA: MINOR FUNGAL PHYLA<br />

Fig 4.8 Thraus<strong>to</strong>chytriales. (a) Thallus of Thraus<strong>to</strong>chytrium sp. growing on a pollen grain sprinkled on<strong>to</strong> seawater. (b) Thalli of<br />

Schizochytrium sp. growing on a pollen grain. (c) Thalli of Schizochytrium sp. growing on agar medium. Note the slime net extending<br />

away from the thalli.<br />

and these can settle on<strong>to</strong> a suitable substratum,<br />

giving rise <strong>to</strong> new thalli (Porter, 1990).<br />

Thus, these zoospores of Thraus<strong>to</strong>chytriales<br />

are mi<strong>to</strong>spores formed following mi<strong>to</strong>sis, in<br />

contrast with those of Labyrinthulales which<br />

are meiospores, i.e. formed by meiosis. Although<br />

thraus<strong>to</strong>chytrid zoospores lack a recognizable<br />

eye-spot, they are pho<strong>to</strong>tropic, reacting <strong>to</strong> light<br />

of blue wavelengths such as that produced by<br />

bioluminescent bacteria (Amon & French, 2004).<br />

Chemotropism has also been described for<br />

thraus<strong>to</strong>chytrid zoospores (Fan et al., 2002), and<br />

both sensual responses may enable zoospores<br />

<strong>to</strong> locate potential food sources.<br />

Thraus<strong>to</strong>chytrids, and especially the genera<br />

Thraus<strong>to</strong>chytrium and Schizochytrium, have recently<br />

attracted attention as producers of polyunsaturated<br />

fatty acids (PUFAs). These are important<br />

as nutrient supplements, and thraus<strong>to</strong>chytrid<br />

oils might eventually be able <strong>to</strong> compete with<br />

fish oils on the market (Yokochi et al., 1998;<br />

Lewis et al., 1999).


5<br />

Straminipila: Oomycota<br />

5.1 <strong>Introduction</strong><br />

The phylum Oomycota, alternatively called<br />

Peronosporomycetes (Dick, 2001a), currently comprises<br />

some 800 1000 species (Kirk et al., 2001).<br />

The Oomycota as a whole have been resolved<br />

as a monophyletic group within the kingdom<br />

Straminipila in recent phylogenetic studies (e.g.<br />

Riethmüller et al., 1999; Hudspeth et al., 2000;<br />

see Fig. 4.2), although considerable rearrangements<br />

are still being performed at the level of<br />

orders and families. A scholarly treatment of the<br />

Oomycota has been published by Dick (2001a)<br />

and will remain the reference work for many<br />

years <strong>to</strong> come. Because of the outstanding significance<br />

of Oomycota, especially in plant pathology,<br />

we give an extended treatment of this group.<br />

5.1.1 The vegetative hypha<br />

Although some members of the Oomycota grow<br />

as sac-like or branched thalli, most of them<br />

produce hyphae forming a mycelium. Oomycota<br />

are now known <strong>to</strong> be the result of convergent<br />

evolution with the true fungi (Eumycota), and<br />

their hyphae differ in certain details. However,<br />

the overall functional similarities are so great<br />

that they provide a persuasive argument for<br />

the fundamental importance of the hypha in<br />

the lifestyle of fungi (Barr, 1992; Carlile, 1995;<br />

Bartnicki-Garcia, 1996). Much physiological work<br />

has been carried out on hyphae of Oomycota (see<br />

Chapter 1), and the results have a direct bearing<br />

on our understanding of the biology of the<br />

Eumycota. Like them, the hyphae of Oomycota<br />

display apical growth and enzyme secretion,<br />

ramify throughout the substratum by branching<br />

<strong>to</strong> form a mycelium, and can show morphogenetic<br />

plasticity by differentiation in<strong>to</strong> specialized<br />

structures such as appressoria or haus<strong>to</strong>ria.<br />

The hyphae of Oomycota are coenocytic, i.e.<br />

they generally do not form cross-walls (septa)<br />

except in old compartments or at the base of<br />

reproductive structures. The cy<strong>to</strong>plasm is generally<br />

coarsely granular and contains vacuoles,<br />

Golgi stacks, mi<strong>to</strong>chondria and diploid nuclei.<br />

The apex is devoid of organelles other than<br />

numerous secre<strong>to</strong>ry vesicles. These are not, as<br />

in the Eumycota, arranged in<strong>to</strong> a Spitzenkörper<br />

because the microvesicles which contain chitin<br />

synthase and make up the Spitzenkörper core are<br />

lacking. This is in line with the general absence,<br />

with a few exceptions, of chitin from the walls of<br />

Oomycota; instead, cellulose, a crystalline b-(1,4)-<br />

glucan, contributes the main fibrous component.<br />

As in the Eumycota, these structural fibres<br />

are cross-linked by branched b-(1,3)- and b-(1,6)-<br />

glucans, although the biochemical properties of<br />

the glucan synthases seem <strong>to</strong> differ fundamentally<br />

between those of Eumycota on the one hand<br />

and those of Oomycota and plants on the other<br />

(Antelo et al., 1998). Other biochemical differences<br />

include the lysine synthetic pathway (DAP<br />

in plants and Oomycota; AAA in true <strong>Fungi</strong>;<br />

see p. 67) and details of sterol metabolism (Nes,<br />

1990; Dick, 2001a).<br />

The mi<strong>to</strong>chondria of Oomycota are indistinguishable<br />

by light microscopy from those of the<br />

Eumycota, but when viewed with the transmission<br />

electron microscope they have tubular


76 STRAMINIPILA: OOMYCOTA<br />

Fig 5.1 Asexual reproductive stages in Saprolegnia.<br />

(a) Auxiliary (primary) zoospore. (b) Principal<br />

(secondary) zoospore. Schematic drawings, based<br />

partly on Dick (2001a).<br />

rather than lamellate cristae (see Fig. 4.1). The<br />

vacuolar system of Oomycota is also unusual in<br />

containing dense-body vesicles or ‘fingerprint<br />

vacuoles’ (see Fig. 5.24b) which consist of deposits<br />

of a phosphorylated b-(1,3)-glucan polymer, mycolaminarin.<br />

Mycolaminarin may serve as a s<strong>to</strong>rage<br />

compound for carbohydrates as well as phosphate<br />

(Hemmes, 1983), and the polyphosphate<br />

s<strong>to</strong>rage deposits which are typically found within<br />

vacuoles of true <strong>Fungi</strong> are absent from vacuoles<br />

of Oomycota (Chilvers et al., 1985). Apart from<br />

that, however, vacuoles of Oomycota share many<br />

features with those of true <strong>Fungi</strong>, including the<br />

membranous continuities which often link adjacent<br />

vacuoles and provide a means of transport<br />

by peristalsis (Rees et al., 1994; see Fig. 1.9).<br />

Cy<strong>to</strong>plasmic glycogen granules, which are one<br />

of the major carbohydrate s<strong>to</strong>rage sites in<br />

Eumycota, are absent from hyphae of Oomycota<br />

(Bartnicki-Garcia & Wang, 1983).<br />

5.1.2 The zoospore<br />

The Oomycota are characterized by motile asexual<br />

spores (zoospores) which are produced in<br />

spherical or elongated zoosporangia. They are<br />

heterokont, possessing one straminipilous and<br />

one whiplash-type flagellum. Two types of zoospore<br />

may be produced and, if so, the auxiliary<br />

zoospore is the first formed. It is grapeseedshaped,<br />

with both flagella inserted apically<br />

(Fig. 5.1a), and it encysts soon after its formation.<br />

Encystment is by withdrawal of the flagella,<br />

so that a tuft of tripartite tubular hairs (TTHs;<br />

see p. 68) is left behind on the surface of the<br />

developing cyst (Dick, 2001b). The cyst germinates<br />

<strong>to</strong> give rise <strong>to</strong> the principal zoospore, which is<br />

by far the more common type and also the more<br />

vigorous swimmer. This typical and readily recognized<br />

oomycete zoospore is uniform in appearance<br />

across the phylum (Lange & Olson, 1983;<br />

Dick, 2001a). In species lacking auxiliary zoospores,<br />

the principal zoospore is usually produced<br />

directly from a sporangium. It is kidney-shaped,<br />

with the flagella inserted laterally in a kine<strong>to</strong>some<br />

boss which in turn is located within the<br />

lateral groove (Fig. 5.1b). Encystment is initiated<br />

by the shedding, rather than withdrawal, of the<br />

flagella; no tufts of TTHs are left on the cyst<br />

surface (Dick, 2001a). Fascinating insights in<strong>to</strong><br />

the cy<strong>to</strong>logy of zoospore encystment have been


INTRODUCTION<br />

77<br />

Fig 5.2 Schematic drawing and terminology of sexual<br />

reproductive organs in the Oomycota. Modified from Dick<br />

(1995).<br />

obtained from several species (see Fig. 5.24). At<br />

the onset of encystment, adhesive and cell wall<br />

material is secreted by the synchronized fusion<br />

of pre-formed s<strong>to</strong>rage vesicles with the zoospore<br />

plasma membrane (Hardham et al., 1991;<br />

Hardham, 1995), thereby providing a rare example<br />

of regulated secretion in fungi. Constitutive<br />

secretion by growing hyphal tips is more commonly<br />

associated with their mode of life.<br />

Some members of the Oomycota have no<br />

motile spore stages but can be readily related<br />

<strong>to</strong> groups still producing them.<br />

5.1.3 Sexual reproduction<br />

The life cycle of the Oomycota is of the<br />

haplomi<strong>to</strong>tic B type, i.e. mi<strong>to</strong>sis occurs only<br />

between karyogamy and meiosis. All vegetative<br />

structures of Oomycota are therefore diploid<br />

(see Figs. 5.3 and 5.19). This is in contrast <strong>to</strong> the<br />

Eumycota in which vegetative nuclei are usually<br />

haploid, the first division after karyogamy being<br />

meiotic. Sexual reproduction in Oomycota is<br />

oogamous, i.e. male and female gametangia are<br />

of different size and shape (Fig. 5.2). Meiosis<br />

occurs in the male antheridia and in the female<br />

oogonia, and is followed by plasmogamy (fusion<br />

between the pro<strong>to</strong>plasts) and karyogamy (fusion<br />

of haploid nuclei). Numerous meioses can occur<br />

synchronously, so that true sexual reproduction<br />

can actually happen within the same pro<strong>to</strong>plast<br />

(Dick, 1990a). Heterothallic species of Oomycota<br />

display relative sexuality, i.e. a strain can produce<br />

antheridia in combination with a second<br />

strain but oogonia when paired against a third<br />

(see pp. 86 and 95). Steroid hormones play<br />

an important role in sexual reproduction (see<br />

Fig. 5.11).<br />

The mature oospore contains three major<br />

pools of s<strong>to</strong>rage compounds (Fig. 5.2; Dick, 1995).<br />

The oospore wall often appears stratified, and<br />

this is due in part <strong>to</strong> a polysaccharide reserve<br />

compartment, the endospore, which is located


78 STRAMINIPILA: OOMYCOTA<br />

Fig 5.3 Life cycle of Saprolegnia.Vegetative hyphae are diploid and coenocytic. Asexual reproduction is by means of diplanetic<br />

(auxiliary and principal) zoospores.The principal zoospore state is polyplanetic. Saprolegnia is homothallic, and sexual reproduction<br />

is initiated by the formation of antheridia and oogonia.For simplicity, only a single nucleus is shown in each of the oospheres and<br />

in the antheridium. Each oogonium contains several oospheres. Karyogamy occurs soon after fertilization of an oosphere by an<br />

antheridial nucleus.The oospore may germinate by means of a germ sporangium (not shown) or a hyphal tip.Open and closed circles<br />

represent haploid nuclei of opposite mating type; diploid nuclei are larger and half-filled. Key events in the life cycle are meiosis (M),<br />

plasmogamy (P) and karyogamy (K).<br />

between the plasma membrane and the outer<br />

spore wall (epispore). Upon germination, the<br />

endospore is thought <strong>to</strong> coat the emerging germ<br />

tube with wall material, and some material may<br />

also be taken up by endocy<strong>to</strong>sis. A large s<strong>to</strong>rage<br />

vacuole inside the oospore pro<strong>to</strong>plast is called<br />

the ooplast. It arises by fusion of dense-body<br />

vesicles and, like them, contains mycolaminarin<br />

and phosphate. Dick (1995, 2001a) speculated<br />

that the ooplast contributes membrane precursor<br />

material during the process of oospore germination.<br />

The third s<strong>to</strong>rage compartment consists<br />

of one or several lipid droplets which provide<br />

the endogenous energy supply required for germination.<br />

Ultrastructural changes during oospore<br />

germination have been described by Beakes<br />

(1981).<br />

5.1.4 Ecology and significance<br />

Oomycota have a major impact on mankind as<br />

pathogens causing plant diseases of epidemic<br />

proportions. Two events have had particularly<br />

far-reaching political and social consequences,<br />

and have shaped and interlinked the young<br />

disciplines of mycology and plant pathology in<br />

the nineteenth century. These were the great<br />

Irish pota<strong>to</strong> famine of 1845 1848 caused by<br />

Phy<strong>to</strong>phthora infestans (Bourke, 1991), and the<br />

occurrence of downy mildew of grapes caused<br />

by Plasmopara viticola (Large, 1940). The former


SAPROLEGNIALES<br />

79<br />

prepared the way for the then revolutionary<br />

theory that fungal infections can be the cause<br />

rather than the consequence of disease, whereas<br />

the latter stimulated research in<strong>to</strong> chemical<br />

control of diseases which directly gave rise <strong>to</strong><br />

the first fungicide, Bordeaux mixture (p. 119;<br />

Large, 1940).<br />

Although all members of Oomycota depend<br />

on moist conditions for the dispersal of their<br />

zoospores, they are cosmopolitan and ubiqui<strong>to</strong>us<br />

even in terrestrial situations. In species adapted<br />

<strong>to</strong> drier habitats, the sporangia often germinate<br />

directly <strong>to</strong> produce a germ tube, with zoospores<br />

released as an alternative germination method<br />

only in the presence of moisture, or lacking<br />

al<strong>to</strong>gether. Oomycota occur in freshwater, the<br />

sea, in the soil and on above-ground plant<br />

organs. Most are obligate aerobes, although<br />

some <strong>to</strong>lerate anaerobic conditions (Emerson &<br />

Natvig, 1981; Voglmayr et al., 1999), and one<br />

species (Aqualinderella fermentans) is obligately<br />

anaerobic and lacks mi<strong>to</strong>chondria (Emerson &<br />

Wes<strong>to</strong>n, 1967). Oomycota live either saprotrophically<br />

on organic material, or they may be<br />

obligate (biotrophic) or facultative (necrotrophic)<br />

parasites of plants. Some can also cause diseases<br />

of animals, such as Aphanomyces astaci which has<br />

all but eliminated European crayfish from many<br />

rivers (p. 94), Saprolegnia spp. which cause serious<br />

infections of farmed fish, especially salmon<br />

(Plate 2a; Dick, 2003), or Pythium insidiosum<br />

causing equine phycomycosis (de Cock et al.,<br />

1987). Yet other Oomycota, notably Lagenidium<br />

giganteum, parasitize insects and may prove<br />

valuable in the biological control of mosqui<strong>to</strong><br />

larvae (Dick, 1998).<br />

5.1.5 Classification<br />

As indicated above, the classification of<br />

Oomycota at the level below the phylum is still<br />

an ongoing process, and it is difficult at present<br />

<strong>to</strong> reconcile the different classification schemes<br />

that are being proposed. Kirk et al. (2001) listed<br />

eight orders in the phylum Oomycota, of which<br />

Dick (2001b) treated six within the class<br />

Peronosporomycetes, his equivalent <strong>to</strong> the<br />

Oomycota, considering the other two of<br />

uncertain affinity (incertae sedes). These groups<br />

are summarized in Table 5.1.<br />

5.2 Saprolegniales<br />

The order Saprolegniales is currently divided<br />

up in<strong>to</strong> two families, the Saprolegniaceae<br />

(e.g. Achlya, Brevilegnia, Dictyuchus, Saprolegnia,<br />

Thraus<strong>to</strong>theca) and Lep<strong>to</strong>legniaceae (Aphanomyces,<br />

Lep<strong>to</strong>legnia, Plec<strong>to</strong>spira), <strong>to</strong>talling 132 species in<br />

about 20 genera (Dick, 2001a; Kirk et al., 2001).<br />

The Saprolegniales are the best-known group of<br />

aquatic fungi, often termed the water moulds.<br />

Members of this group are abundant in wet soils,<br />

lake margins and freshwater, mainly as saprotrophs<br />

on plant and animal debris. Whilst some<br />

Saprolegniales occur in brackish water, most<br />

are in<strong>to</strong>lerant of it and thrive best in freshwater.<br />

A few species of Saprolegnia and Achlya are economically<br />

important as parasites of fish and<br />

their eggs (Willoughby, 1994). Aphanomyces<br />

euteiches causes a root rot of peas and some<br />

other plants, whilst A. astaci is a serious parasite<br />

of the European crayfish Astacus (Alderman et al.,<br />

1990). Algae, fungi, rotifers and copepods may<br />

also be parasitized by members of the group, and<br />

occasional epidemics of disease among zooplank<strong>to</strong>n<br />

have been reported.<br />

Members of the Saprolegniales are characterized<br />

by coarse, stiff hyphae which branch<br />

<strong>to</strong> produce a typically fast-growing mycelium.<br />

The hyphae of Saprolegniales are coenocytic,<br />

containing a peripheral layer of cy<strong>to</strong>plasm<br />

surrounding a continuous central vacuole.<br />

Cy<strong>to</strong>plasmic streaming is readily observed in<br />

the peripheral cy<strong>to</strong>plasm. Numerous nuclei are<br />

present. Mi<strong>to</strong>tic division is associated with the<br />

replication of paired centrioles and the development<br />

of an intranuclear mi<strong>to</strong>tic spindle; the<br />

nuclear membrane remains intact throughout<br />

division (Dick, 1995). Filamen<strong>to</strong>us mi<strong>to</strong>chondria<br />

and lipid droplets can also be observed in<br />

vegetative hyphae. The mi<strong>to</strong>chondria are orientated<br />

parallel <strong>to</strong> the long axis of the hypha and<br />

are sufficiently large <strong>to</strong> be seen in cy<strong>to</strong>plasmic<br />

streaming in living material. Important physiological<br />

work has been carried out on the


80 STRAMINIPILA: OOMYCOTA<br />

Table 5.1. Summary of the most important groups of Oomycota and their characteristic features. Only the<br />

last four groups are considered further in this book. Based on information provided by Dick (2001a,b) and<br />

Kirk et al. (2001).<br />

Order<br />

Number<br />

of species<br />

Thallus and<br />

reproduction<br />

Ecology<br />

Myzocytiopsidales<br />

(incertae sedes)<br />

Olpidiopsidales<br />

(incertae sedes)<br />

74 Holocarpic, later<br />

coralloid or breaking<br />

up in<strong>to</strong> segments.<br />

Zoospores, oospores.<br />

21 Holocarpic, becoming<br />

converted in<strong>to</strong> a<br />

sporangium. Zoospores,<br />

oospores.<br />

Rhipidiales 12 Eucarpic with rhizoids.<br />

Zoospores, oospores.<br />

Lep<strong>to</strong>mitales 25 Constricted hyphae<br />

producing sporangia.<br />

Zoospores, oospores.<br />

Saprolegniales<br />

(see Section 5.2)<br />

Pythiales<br />

(see Section 5.3)<br />

Peronosporales<br />

(see Section 5.4)<br />

Sclerosporaceae<br />

(see Section 5.5)<br />

132 Mycelium of wide s<strong>to</strong>ut<br />

hyphae. Zoospores,<br />

oospores.<br />

4200 Mycelium of relatively<br />

narrow hyphae.<br />

Zoospores, oospores.<br />

252 Intercellular mycelium with<br />

haus<strong>to</strong>ria.Differentiated<br />

sporangiophores.<br />

Zoospores or‘conidia’,<br />

oospores.<br />

22 Mycelium of very narrow<br />

hyphae.Differentiated<br />

sporangiophores.<br />

Zoospores or‘conidia’,<br />

oospores.<br />

Parasites of invertebrates or algae.<br />

Biotrophic parasites of Oomycota,<br />

Chytridiomycota and algae.<br />

Freshwater saprotrophs, facultatively<br />

or obligately anaerobic.<br />

Freshwater saprotrophs or parasites<br />

of animals.<br />

Saprotrophs or necrotrophic pathogens<br />

of animals, plants and other organisms.<br />

Saprotrophs or pathogens<br />

(often necrotrophic) of plants, fungi and<br />

animals.<br />

Biotrophic plant pathogens, causing<br />

downy mildews and other diseases.<br />

Biotrophic pathogens of grasses, causing<br />

downy mildews.<br />

For thallus terminology, see Fig.6.1.<br />

mechanisms of hyphal polarity and growth<br />

regulation in Achlya and Saprolegnia (see Heath,<br />

1995b; Hyde & Heath, 1997; Heath & Steinberg,<br />

1999). Like other Oomycota but in contrast <strong>to</strong> the<br />

Eumycota (Pfyffer et al., 1986; Rast & Pfyffer,<br />

1989), these fungi are unable <strong>to</strong> synthesize<br />

compatible osmotically active solutes such as<br />

glycerol, manni<strong>to</strong>l and other polyols <strong>to</strong> maintain<br />

their intrahyphal turgor pressure against fluctuating<br />

external conditions. Under conditions of<br />

water stress, the turgor pressure in hyphae of<br />

Achlya and Saprolegnia approaches zero, yet hyphal<br />

growth can still occur at least under labora<strong>to</strong>ry<br />

conditions because of the enhanced secretion of


SAPROLEGNIALES<br />

81<br />

cell wall-softening enzymes and the role of the<br />

cy<strong>to</strong>skele<strong>to</strong>n in pushing forward the growing tip<br />

(see pp. 6 9; Money & Harold, 1992, 1993; Money,<br />

1997; Money & Hill, 1997).<br />

The Saprolegniales are the only order within<br />

the Oomycota <strong>to</strong> produce both auxiliary and<br />

principal zoospores, although both forms are not<br />

produced in all genera. The production of two<br />

distinct motile stages is termed diplanetism.<br />

It has also been called dimorphism, but this<br />

term has several different meanings and is best<br />

avoided in the current context. Depending on<br />

environmental conditions, the cysts of principal<br />

zoospores may germinate either by means of a<br />

germ tube developing in<strong>to</strong> a hypha or by the<br />

emergence of a new principal zoospore. The<br />

repetition of the same type of motile spore is<br />

called polyplanetism.<br />

Sexual reproduction in the Saprolegniales is<br />

oogamous, with a large, usually spherical oogonium<br />

containing one or several oospheres.<br />

Antheridial branches apply themselves <strong>to</strong> the<br />

wall of the oogonium and penetrate the wall<br />

by fertilization tubes through which a single<br />

nucleus is introduced in<strong>to</strong> each oosphere. A<br />

feature of many Saprolegniales, especially when<br />

grown in culture, is the formation of thick-walled<br />

enlarged terminal or intercalary portions of<br />

hyphae which become packed with dense cy<strong>to</strong>plasm<br />

and are cut off from the rest of the<br />

mycelium by septa. These structures, which may<br />

occur singly or in chains (see Fig. 5.6g), are termed<br />

gemmae or chlamydospores, and their formation<br />

can be induced by manipulating the culture conditions.<br />

Morphologically less distinct but otherwise<br />

similar structures are frequently found in<br />

old cultures. Although it is known that chlamydospores<br />

cannot survive desiccation or prolonged<br />

freezing, they remain viable for long periods in<br />

less extreme conditions. They may function as<br />

female gametangia or as zoosporangia, but more<br />

frequently they germinate by means of a germ<br />

tube. Another feature of old cultures is the<br />

fragmentation of cylindrical pieces of mycelium<br />

cut off at each end by a septum.<br />

Members of the Saprolegniales can be isolated<br />

readily from water, mud and soil by floating split<br />

boiled hemp seeds or dead house flies in dishes<br />

containing pond water, or by covering soil<br />

samples or waterlogged twigs with water<br />

(Stevens, 1974; Dick, 1990a). Within about 4 days<br />

the fungi can be recognized by their stiff,<br />

radiating, coarse hyphae bearing terminal sporangia,<br />

and cultures can be prepared by transferring<br />

hyphal tips or zoospores <strong>to</strong> cornmeal agar<br />

or other suitable media. The most commonly<br />

encountered genera are Achlya, Dictyuchus,<br />

Saprolegnia, Thraus<strong>to</strong>theca and Aphanomyces. With<br />

the exception of a few obligately parasitic species,<br />

most of the Saprolegniales will grow readily in<br />

pure culture even on chemically defined media,<br />

and extensive studies of their nutritional physiology<br />

have been undertaken (summarized by<br />

Cantino, 1955; Gleason, 1976; Jennings, 1995).<br />

Most species examined have no requirement for<br />

vitamins. Organic forms of sulphur such as<br />

cysteine, cystine, glutathione and methionine<br />

are preferred, and most species are unable <strong>to</strong><br />

reduce sulphate. Organic nitrogen sources such<br />

as amino acids, pep<strong>to</strong>ne and casein are preferred<br />

<strong>to</strong> inorganic sources. Ammonium is widely<br />

utilized, but nitrate is not. Glucose is the most<br />

widely utilized carbon source, but many species<br />

also degrade mal<strong>to</strong>se, starch and glycogen. In<br />

liquid culture, Saprolegnia can be maintained in<br />

the vegetative state indefinitely if supplied with<br />

organic nutrients in the form of broth. When the<br />

nutrients are replaced by water, the hyphal tips<br />

quickly develop in<strong>to</strong> zoosporangia. The formation<br />

of sexual organs can similarly be affected by<br />

manipulating the external conditions in some<br />

species, and the concentration of salts in the<br />

medium may play a decisive role (Barksdale,<br />

1962; Davey & Papavizas, 1962).<br />

5.2.1 Saprolegnia (Saprolegniaceae)<br />

Species of Saprolegnia are common in soil and in<br />

freshwater as saprotrophs on plant and animal<br />

remains. A few species such as S. parasitica and<br />

S. polymorpha cause disease in fish and their eggs<br />

(Plate 2a). Salmonid fish are particularly affected,<br />

and the disease can cause significant damage<br />

in fish farms around the world (Willoughby,<br />

1994, 1998a). Control by fungicides is difficult<br />

but possible (Willoughby & Roberts, 1992). The<br />

disease is also seen in wild salmon and other fish<br />

(Söderhäll et al., 1991; Bly et al., 1992). Pathogenic


82 STRAMINIPILA: OOMYCOTA<br />

strains or species may be closely related <strong>to</strong> nonpathogenic<br />

ones but can be distinguished by<br />

physiological characteristics, DNA sequencing<br />

(Yuasa & Hatai, 1996) and the length of the<br />

‘boat hook’ appendages on the cysts of principal<br />

zoospores (Figs. 5.5b,c; Beakes, 1983; Burr &<br />

Beakes, 1994).<br />

The life cycle of Saprolegnia is summarized in<br />

Fig. 5.3. A monographic treatment of the genus<br />

has been published by Seymour (1970).<br />

Asexual reproduction in Saprolegnia<br />

Sporangia of Saprolegnia develop when a hyphal<br />

tip, which is pointed in the vegetative condition,<br />

swells, rounds off and becomes club-shaped. It<br />

accumulates denser cy<strong>to</strong>plasm around the<br />

vacuole which remains clearly visible. A septum<br />

develops at the sporangial base and it is at first<br />

straight or convex with respect <strong>to</strong> the sporangium,<br />

i.e. it bulges in<strong>to</strong> it (Figs. 5.4c,d). The<br />

sporangium contains numerous nuclei, and<br />

Fig 5.4 Saprolegnia. (a) Apex of vegetative hypha. (b d) Stages in the development of zoosporangia. (e) Release of zoospores.<br />

(f) Proliferation of zoosporangium. A second zoosporangium is developing within the empty one. (g) Auxiliary zoospore (first<br />

motile stage). (h) Cyst formed at the end of the first motile stage (auxiliary cyst). (i,j) Germination of auxiliary cyst <strong>to</strong> release a<br />

second motile stage (principal zoospores).These have the typical reniform shape. (k m) Principal zoospores. (n) Principal zoospore<br />

at the moment of encystment. Note the shed flagellum. (o) Principal cyst. (p) Principal cyst germinating by means of a germ tube.<br />

(a f) <strong>to</strong> same scale; (g p) <strong>to</strong> same scale. Note that the straminipilous flagellum cannot be distinguished from the whiplash flagellum<br />

at the magnification chosen.


SAPROLEGNIALES<br />

83<br />

cleavage furrows separate the cy<strong>to</strong>plasm in<strong>to</strong><br />

uninucleate pieces, each of which differentiates<br />

in<strong>to</strong> an auxiliary zoospore. As the zoospores are<br />

cleaved, the central vacuole disappears. The tip<br />

of the cylindrical sporangium contains clearer<br />

cy<strong>to</strong>plasm and a flattened protuberance, the<br />

papilla, develops at the apex. As the sporangium<br />

ripens and the zoospores become fully<br />

differentiated, they show limited movement<br />

and change of shape (Figs. 5.4b d). Shortly<br />

before discharge, there is evidence of a buildup<br />

of turgor pressure within the sporangium<br />

because the basal septum becomes concave,<br />

i.e. it is bent <strong>to</strong>wards the lumen of the hypha<br />

beneath the sporangium. After cleavage, the<br />

positive turgor pressure is lost concomitantly<br />

with the loss of the sporangial plasma membrane<br />

which becomes part of the zoospore<br />

membranes, and the septum again bulges in<strong>to</strong><br />

the sporangium while the zoospores become fully<br />

differentiated. The sporangium undergoes a<br />

slight change of shape at this time and the<br />

sporangium wall breaks down at the papilla. The<br />

spores are released quickly, many zoospores<br />

escaping in a few seconds and moving as a<br />

column through the opening. Osmotic phenomena<br />

have been invoked <strong>to</strong> explain the rapidity<br />

of discharge, and the osmotically active substances<br />

must be large enough <strong>to</strong> be contained<br />

by the sporangial wall. Mycolaminarin, released<br />

from the central vacuole during zoospore differentiation,<br />

is the likely solute (Money &<br />

Webster, 1989). The whole process of sporangium<br />

differentiation takes about 90 min. The zoospores<br />

leave the sporangium backwards, with the blunt<br />

posterior end emerging first. The size of the<br />

zoospore is sometimes greater than the diameter<br />

of the sporangial opening so that the zoospores<br />

are squeezed through it. An occasional zoospore<br />

may be left behind, swimming about in the<br />

empty zoosporangium for a while before making<br />

its exit. Zoospores in partially empty sporangia<br />

orientate themselves in a linear fashion along the<br />

central axis of the sporangium.<br />

A characteristic feature of Saprolegnia is that,<br />

following the discharge of a zoosporangium,<br />

growth is renewed from the septum at its base so<br />

that a new apex develops inside the old sporangial<br />

wall by internal proliferation. This in<br />

turn may develop in<strong>to</strong> a zoosporangium, discharging<br />

its spores through the old pore (Fig. 5.4f).<br />

The process may be repeated so that several<br />

empty zoosporangial walls may be found inside,<br />

or partially inside, each other.<br />

Upon release, the auxiliary zoospores slowly<br />

revolve and eventually swim somewhat sluggishly<br />

with the pointed end directed forwards.<br />

They are grapeseed- or pear-shaped (‘Conference’<br />

pear; Dick, 2001a) and bear two apically attached<br />

flagella (see Figs. 5.1a, 5.4g). Each zoospore<br />

also contains a diploid nucleus, mi<strong>to</strong>chondria,<br />

a contractile vacuole and numerous vesicles<br />

(Holloway & Heath, 1977a,b). The zoospores<br />

from a single sporangium show variation in<br />

their period of motility, the majority encysting<br />

within about a minute, but some remaining<br />

motile for over an hour. The zoospore then<br />

withdraws its flagella and encysts, i.e. the<br />

cy<strong>to</strong>plasm becomes surrounded by a distinct<br />

wall which is produced from pre-formed material<br />

s<strong>to</strong>red in the cy<strong>to</strong>plasmic vesicles. Only the<br />

axonemes of the flagella are withdrawn, leaving<br />

the TTHs of the straminipilous flagellum at<br />

the surface of the cyst (see Fig. 5.5a). Following<br />

a period of rest (2 3h in S. dioica), the cyst<br />

germinates <strong>to</strong> release a further zoospore, the<br />

principal zoospore (Figs. 5.4i,j). This differs in<br />

shape from the auxiliary zoospore in being beanshaped,<br />

with the two flagella inserted laterally<br />

in a shallow groove running down one side of<br />

the zoospore (Fig. 5.1b). The principal zoospore<br />

may swim vigorously for several hours before<br />

encysting. Salvin (1941) compared the rates of<br />

movement of auxiliary and principal zoospores<br />

in Saprolegnia and found that the latter swam<br />

about three times more rapidly. The probable<br />

reason for this is that the lateral insertion of<br />

both flagella allows the straminipilous flagellum<br />

<strong>to</strong> point forward and the whiplash one <strong>to</strong> point<br />

backward, thereby improving the propulsion<br />

relative <strong>to</strong> the apical insertion in which both<br />

flagella point forward.<br />

Movement of principal zoospores is chemotactic<br />

and zoospores can be stimulated <strong>to</strong><br />

aggregate on parts of animal bodies such as the<br />

leg of a fly, or the surface of a fish (Fischer &<br />

Werner, 1958; Willoughby & Pickering, 1977).<br />

When principal zoospores encyst, they shed


84 STRAMINIPILA: OOMYCOTA<br />

Fig 5.5 Surface features of Saprolegnia. (a) Detail of<br />

an auxiliary zoospore cyst of S. parasitica showing<br />

the tuft of TTHs (mt) at the point where the<br />

straminipilous flagellum was withdrawn. (b) Surface of<br />

a principal zoospore cyst of S. parasitica; the long boat<br />

hook spines are arranged in fascicles. (c) Surface of a<br />

principal zoospore cyst of S. hypogyna with discrete<br />

boat hooks of intermediate length. All bars ¼ 2 mm.<br />

All images kindly provided by M.W. Dick and I.C.<br />

Hallett; (b) reprinted from Hallett and Dick (1986),<br />

with permission from Elsevier.<br />

rather than withdraw their flagella. The first step<br />

in encystment is the fusion of vesicles called<br />

K-bodies with the plasma membrane. These are<br />

so called because they are located near the<br />

kine<strong>to</strong>some. The material they secrete is involved<br />

in attachment of the zoospore <strong>to</strong> a substratum,<br />

which occurs in the region of the groove near the<br />

flagellar bases, designated the ventral region<br />

(Lehnen & Powell, 1989). The cyst wall and preformed<br />

boat hook spines are secreted by fusion of<br />

encystment vesicles with the plasma membrane<br />

(Beakes, 1987; Burr & Beakes, 1994). The length<br />

and arrangement of spines on the surface of a<br />

mature principal cyst are characteristic features<br />

of individual species (Figs. 5.5b,c). They probably<br />

mediate attachment of the cyst <strong>to</strong> the host, and<br />

pathogenic isolates of Saprolegnia have much<br />

longer spines than saprotrophic ones (Burr &<br />

Beakes, 1994). Alternatively, the boat hooks may<br />

mediate attachment <strong>to</strong> the water meniscus.


SAPROLEGNIALES<br />

85<br />

Either way, attachment must be very effective<br />

because trout or char, placed in a water bath with<br />

principal zoospores of S. parasitica for 10 min and<br />

followed by 1 h in clean water, had an extremely<br />

high concentration of cysts attached <strong>to</strong> the skin<br />

(Willoughby & Pickering, 1977).<br />

Principal zoospore cysts can germinate either<br />

by means of a germ tube (Fig. 5.4p) or by<br />

releasing a further principal zoospore which in<br />

turn may germinate directly or by releasing yet<br />

another motile stage. Saprolegnia is therefore<br />

polyplanetic. The auxiliary and principal zoospores,<br />

as well as the cysts they form, differ<br />

morphologically from each other, i.e. they are<br />

diplanetic.<br />

Sexual reproduction in Saprolegnia<br />

All members of the genus Saprolegnia characterized<br />

<strong>to</strong> date are homothallic, i.e. a culture<br />

derived from a single zoospore will give rise <strong>to</strong><br />

a mycelium forming both oogonia and antheridia.<br />

In contrast, Achlya also contains heterothallic<br />

species in which sexual reproduction occurs<br />

only when two different strains are juxtaposed,<br />

one forming oogonia, the other antheridia (see<br />

Fig. 5.10).<br />

Sexual reproduction follows a similar course<br />

in all members of the Saprolegniales. Oogonia<br />

containing one or several eggs are fertilized by<br />

antheridial branches. Fertilization is accomplished<br />

by the penetration of fertilization tubes<br />

in<strong>to</strong> the oogonium. In some species, ripe oogonia<br />

are found without antheridia associated with<br />

them (Fig. 5.6f); this could be due either <strong>to</strong> the<br />

fusion of two haploid nuclei from adjacent<br />

meiotic events in a single oogonium (apomixis)<br />

or the formation of an oospore around a diploid<br />

nucleus that never underwent meiosis (parthenogenesis).<br />

Both processes are impossible <strong>to</strong><br />

distinguish without detailed cy<strong>to</strong>logical evidence<br />

(Dick, 2001a). The typical arrangement of oogonia<br />

and antheridia in Saprolegnia is shown in<br />

Fig. 5.6. Antheridial branches arising from the<br />

stalk of the oogonium or the same hypha as the<br />

oogonium are said <strong>to</strong> be monoclinous whereas<br />

they are diclinous if they originate from different<br />

hyphae.<br />

Fig 5.6 Saprolegnia li<strong>to</strong>ralis.(a d) Stages in the development of oogonia. (c) Oogonium showing furrowed cy<strong>to</strong>plasm indicative<br />

of centrifugal cleavage. (d) Outlines of two oospheres become visible. (e) Oogonium with two mature oospores. (f) Intercalary<br />

oogonium lacking antheridia.The oospores have developed by apomixis or parthogenesis. (g) Chain of chlamydospores.


86 STRAMINIPILA: OOMYCOTA<br />

The oogonial initial is multinucleate, and<br />

nuclear divisions continue as it enlarges.<br />

Eventually some of the nuclei degenerate, leaving<br />

only those nuclei which are included in the<br />

oospheres. From the central vacuole within the<br />

oogonium, cleavage furrows radiate outwards <strong>to</strong><br />

divide the cy<strong>to</strong>plasm in<strong>to</strong> uninucleate portions<br />

which round off <strong>to</strong> form oospheres. Oogonium<br />

differentiation is thus centrifugal, which is<br />

typical of the Saprolegniales. Cleavage of the<br />

oospheres from the cy<strong>to</strong>plasm is brought about<br />

by the coalescence of dense body vesicles which<br />

finally fuse with the plasma membrane of the<br />

oogonium so that the oospheres tumble in<strong>to</strong> the<br />

centre of the oogonium (Dick, 2001a). The entire<br />

mass of cy<strong>to</strong>plasm within the oogonium is used<br />

up in the formation of oospheres and there is<br />

no residual cy<strong>to</strong>plasm (periplasm) as in the<br />

Peronosporales. The wall of the oogonium is<br />

often uniformly thick, but in some species it<br />

shows thin areas or pits through which fertilization<br />

tubes may enter (Fig. 5.6e). A septum at<br />

the base of the oogonium cuts it off from the<br />

subtending hypha.<br />

The antheridia are also multinucleate. The<br />

antheridial branch grows <strong>to</strong>wards the oogonium<br />

and attaches itself <strong>to</strong> the oogonial wall. The tip of<br />

the antheridial branch is cut off by a septum,<br />

and the resulting antheridium puts out a fertilization<br />

tube which penetrates the oogonial wall<br />

and may branch within the oogonium. After the<br />

tube has penetrated an oosphere wall, a male<br />

nucleus eventually fuses with the single oosphere<br />

nucleus. The fertilized oosphere (oospore) undergoes<br />

a series of changes described by Beakes<br />

and Gay (1978a,b). The wall of the oospore<br />

thickens and oil globules become obvious.<br />

Mature oospores contain a membrane-bound<br />

vacuole-like body, the ooplast, surrounded by<br />

cy<strong>to</strong>plasm containing various organelles, with<br />

lipid droplets particularly prominently visible.<br />

In Saprolegnia, the ooplast contains particles in<br />

Brownian motion. The position of the ooplast in<br />

the oospore is used for species identification, and<br />

four types of oospore have been distinguished<br />

(Fig. 5.7; Seymour, 1970; Howard, 1971). Centric<br />

oospores have a central ooplast surrounded by<br />

one or two peripheral layers of small lipid<br />

droplets (e.g. S. hypogyna, S. ferax). Subcentric<br />

oospores have several layers of small lipid<br />

droplets on one side of the ooplast and only one<br />

layer or none at all on the other (e.g. S. unispora,<br />

S. terrestris). In subeccentric oospores, the small<br />

lipid droplets have fused in<strong>to</strong> several large ones<br />

all grouped <strong>to</strong> one side, with the ooplast contacting<br />

the plasma membrane on the opposite side<br />

(e.g. S. eccentrica). The eccentric type (found, for<br />

example, in S. anisospora) is similar <strong>to</strong> the<br />

subeccentric type except that there is only one<br />

very large lipid drop. These descriptive terms are<br />

also used for many other species of Oomycota.<br />

5.2.2 Achlya (Saprolegniaceae)<br />

Phylogenetic analyses have shown that the<br />

genera Achlya and Saprolegnia as well as minor<br />

genera of the Saprolegniales are closely related<br />

<strong>to</strong> each other, with possible overlaps which may<br />

necessitate the re-assignment of some species in<br />

future (Riethmüller et al., 1999; Leclerc et al.,<br />

2000; Dick, 2001a). Morphologically and ecologically,<br />

Achlya and Saprolegnia also share several key<br />

features. Both are common in soil and in waterlogged<br />

plant debris such as twigs, and certain<br />

species are pathogens of fish (Willoughby, 1994;<br />

Kitancharoen et al., 1995). Unlike Saprolegnia,<br />

some species of Achlya are heterothallic, but<br />

their life cycle is otherwise similar <strong>to</strong> that of<br />

Saprolegnia given in Fig. 5.3. Heterothallic strains<br />

of Achlya have been the subject of classical<br />

Fig 5.7 Possible arrangements of the<br />

ooplast (shaded organelle) and lipid<br />

droplets (empty circles or ellipses) in<br />

oospores of Saprolegnia. (a) Centric.<br />

(b) Subcentric. (c) Subeccentric.<br />

(d) Eccentric.


SAPROLEGNIALES<br />

87<br />

studies on the nature of mating hormones<br />

(pheromones); additionally, more recent work<br />

has focused on zoospore release. Both aspects are<br />

described below.<br />

Asexual reproduction in Achlya<br />

The development of zoosporangia in Achlya is<br />

similar in all aspects <strong>to</strong> that in Saprolegnia but<br />

has been better researched. The central vacuole<br />

in the developing cylindrical sporangium is<br />

typical of the Saprolegniales and originates<br />

from the fusion of dense body vesicles containing<br />

mycolaminarin. The centrifugal cleavage of<br />

cy<strong>to</strong>plasm from the vacuole <strong>to</strong>wards the plasma<br />

membrane, and the partitioning of individual<br />

spores, are controlled mainly by the actin<br />

cy<strong>to</strong>skele<strong>to</strong>n (Heath & Harold, 1992). In the<br />

Pythiales, vital roles of microtubules in the<br />

organization of differentiating cy<strong>to</strong>plasm have<br />

been described (see p. 102), and microtubules<br />

may have similar but as yet undescribed functions<br />

in the Saprolegniales. As the plasma<br />

membrane of the Achlya zoosporangium is<br />

breached, the zoosporangial volume decreases<br />

by about 10% due <strong>to</strong> the loss of turgor pressure.<br />

Since the membranes of the vacuole contribute<br />

<strong>to</strong> the zoospore plasma membrane, the vacuolar<br />

contents of water-soluble mycolaminarins (b-1,3-<br />

glucans) are released in<strong>to</strong> the sporangium. These<br />

molecules are osmotically active but are <strong>to</strong>o<br />

large <strong>to</strong> diffuse through the sporangial wall, thus<br />

causing the osmotic inward movement of water<br />

in<strong>to</strong> the sporangium, which in turn pressurizes<br />

the sporangium and drives the rapid discharge of<br />

the auxiliary zoospores (Money & Webster, 1985,<br />

1988; Money et al., 1988).<br />

On discharge, the zoospores do not swim<br />

away but cluster in a hollow ball at the mouth of<br />

the zoosporangium and encyst there (Fig. 5.8a).<br />

In fact, it is doubtful whether the term ‘zoospore’<br />

is al<strong>to</strong>gether appropriate as functional flagella<br />

are probably not formed. Partial fragmentation<br />

of the cyst ball frequently occurs and may have<br />

ecological significance in the dispersal of cysts<br />

prior <strong>to</strong> the release of principal zoospores. Unlike<br />

certain species of Saprolegnia, Achlya cysts are<br />

normally found at the bot<strong>to</strong>m of culture dishes,<br />

and presumably also at the water/bot<strong>to</strong>m sediment<br />

interface in natural environments. Cysts of<br />

Fig 5.8 Achlya colorata. (a) Zoosporangium showing a clump<br />

of primary cysts at the mouth.Note the lateral proliferation of<br />

the hypha from beneath the old sporangium. (b) Full and empty<br />

auxiliary cysts. (c) Stages in the release of principal zoospores<br />

from an auxiliary cyst. (d) Principal zoospores. (e) Principal<br />

cyst. (f) Principal cyst germinating by means of a germ tube.<br />

A. klebsiana may remain viable for at least two<br />

months when s<strong>to</strong>red aseptically at 5°C (Reischer,<br />

1951). However, most auxiliary cysts remain at<br />

the mouth of the sporangium for a few hours<br />

and then each cyst releases a principal zoospore<br />

through a small pore (Figs. 5.8b,c). After a period<br />

of swimming, principal zoospores encyst, and<br />

principal cysts germinate either by a germ tube<br />

or by releasing another principal zoospore. When<br />

the zoosporangium of Achlya has released its


88 STRAMINIPILA: OOMYCOTA<br />

Fig 5.9 Achlya colorata.(a d) Stages<br />

in the development of oogonia.<br />

(e) Six-month-old oospores<br />

germinating after 40 h in charcoal<br />

water.<br />

zoospores, growth is usually renewed laterally by<br />

the outgrowth of a new hyphal apex just beneath<br />

the first sporangium (Fig. 5.8a), rather than by<br />

internal proliferation.<br />

Sexual reproduction in Achlya<br />

Some species of Achlya are homothallic (Fig. 5.9)<br />

whereas others are heterothallic (Fig. 5.10).<br />

Achlya colorata, a homothallic species common<br />

in Britain, has oogonial walls which develop<br />

blunt, rounded projections so that the oogonium<br />

appears somewhat spiny (Fig. 5.9d). Otherwise,<br />

the process of sexual reproduction is similar <strong>to</strong><br />

that of Saprolegnia li<strong>to</strong>ralis (Fig. 5.6). Germination<br />

of oospores is often difficult <strong>to</strong> achieve with<br />

Oomycota, but can be stimulated in A. colorata by<br />

transferring mature oospores <strong>to</strong> freshly distilled<br />

water (preferably after shaking with charcoal<br />

and filtering). Germination occurs by means of a<br />

germ tube which grows out from the oospore<br />

through the oogonial wall. Here it may continue<br />

growth as a mycelium (Fig. 5.9e) or may give rise<br />

<strong>to</strong> a sporangium.<br />

The study of heterothallic species of Achlya by<br />

John R. Raper quickly revealed that the formation<br />

of oogonia and antheridia by compatible<br />

strains must be under hormonal control (Raper,<br />

1939, 1957). A particularly readable account of<br />

the classical series of experiments leading <strong>to</strong> the<br />

discovery of the steroid sex hormone, antheridiol<br />

(Fig. 5.11b), has been given by Carlile (1996b).<br />

Several reviews of the broader role of hormones<br />

in fungal reproduction have appeared recently<br />

(Gooday & Adams, 1992; Elliott, 1994). If isolates<br />

of Achlya bisexualis, A. ambisexualis or A. heterosexualis<br />

made from water or mud are grown singly


SAPROLEGNIALES<br />

89<br />

Fig 5.10 Achlya ambisexualis.<br />

(a) Male and female mycelia grown<br />

on hemp seeds and placed <strong>to</strong>gether<br />

in water for 4 days. Note the<br />

formation of antheridial branches<br />

on the male and oogonial branches<br />

on the female. (b) Fertilization,<br />

showing the diclinous origin of the<br />

antheridial branch.<br />

Fig 5.11 Sterols from Achlya spp.Fucosterol (a) is one of the<br />

most abundant sterols in Oomycota and precursor <strong>to</strong> the sex<br />

hormones antheridiol (b) and oogoniol (c).<br />

on hemp seed in water, reproduction is entirely<br />

asexual, but when certain of the isolates are<br />

grown <strong>to</strong>gether in the same dish, it becomes<br />

apparent within 2 3 days that one strain is<br />

forming oogonia, and the other antheridia. The<br />

development of oogonia and antheridia occurs<br />

even when the two strains are held apart in the<br />

water or separated by a cellophane membrane or<br />

by agar. This suggests that one or more diffusible<br />

substances are responsible for the phenomenon.<br />

As compatible colonies approach each other, the<br />

first observable reaction is the production of fine<br />

lateral branches behind the advancing tips of the<br />

male hyphae. These are antheridial branches.<br />

By growing male (antheridial) strains in water<br />

in which a female (oogonial) strain had been<br />

grown previously, Raper (1939) showed that<br />

the vegetative female mycelium was capable<br />

of initiating the development of antheridial<br />

branches on the male. The reverse experiment<br />

showed no effect on female colonies in medium<br />

in which undifferentiated male colonies had<br />

been grown. The role of the vegetative female<br />

colony as initia<strong>to</strong>r of the sequence of events<br />

leading <strong>to</strong> sexual reproduction was confirmed by<br />

ingenious experiments in micro-aquaria consisting<br />

of several consecutive chambers through<br />

which water flowed by means of small siphons.<br />

Male and female colonies were placed alternately<br />

in successive chambers, so that water from a<br />

male colony would flow over a female colony and<br />

so on. If a female colony was placed in the first<br />

chamber, the male colony in the second chamber<br />

reacted by developing antheridial hyphae. If,<br />

however, a male colony was placed in the first<br />

chamber, the male colony in the third chamber


90 STRAMINIPILA: OOMYCOTA<br />

was the first <strong>to</strong> react. Raper (1939) postulated that<br />

the development of the antheridial branches was<br />

in response <strong>to</strong> a hormone, termed Hormone A,<br />

secreted by vegetative female colonies. By further<br />

experiments of this kind, he showed that the later<br />

steps in the sexual process were also regulated<br />

by means of diffusible substances. He postulated<br />

that the antheridial branches secreted a second<br />

substance, Hormone B, which resulted in the<br />

formation of oogonial initials on the female<br />

colony. The oogonial initials in their turn<br />

secreted a further substance called Hormone C,<br />

which stimulated the antheridial initials <strong>to</strong> grow<br />

<strong>to</strong>wards the oogonial initials and also resulted<br />

in the antheridia being delimited. Having made<br />

contact with the oogonial initials, the antheridial<br />

branches secreted Hormone D which resulted<br />

in the formation of a septum cutting off the<br />

oogonium from its stalk, and in the formation<br />

of oospheres. The original scheme (Table 5.2)<br />

therefore implicated four hormones, but confusion<br />

arose subsequently because the effect of<br />

Hormone A can be modulated by amino acids<br />

and other metabolites released from the hemp<br />

seeds (Barksdale, 1970; Schreurs et al., 1989).<br />

Since Hormone A is active at extremely low<br />

concentrations of 2 10 11 M (Barksdale, 1969),<br />

purification of this substance was extremely challenging,<br />

and 6000 l of culture fluid had <strong>to</strong> be<br />

extracted <strong>to</strong> obtain 20 mg crystalline Hormone A<br />

(Barksdale, 1967). It was eventually identified as<br />

the steroid antheridiol (Fig. 5.11b). Soon after, the<br />

structure of Hormone B was elucidated and<br />

found also <strong>to</strong> be a steroid, oogoniol (Fig. 5.11c),<br />

which is, in fact, present as three chemically<br />

closely related forms (McMorris et al., 1975). The<br />

effect postulated by Raper (1939) <strong>to</strong> be due <strong>to</strong><br />

Hormone C is now thought <strong>to</strong> be mediated by<br />

antheridiol activity, whereas Hormone D may not<br />

exist (Carlile, 1996b). Both antheridiol and the<br />

oogoniols are derived from fucosterol (Fig. 5.11a),<br />

the principal sterol in Achlya (see Elliott, 1994).<br />

The physiological roles of antheridiol and<br />

the oogoniols are several-fold and include induction<br />

or suppression of sexuality (Thomas &<br />

McMorris, 1987), directional growth of gametangial<br />

tips (McMorris, 1978), and stimulation of<br />

the production of cell wall-softening enzymes<br />

(especially cellulase) at points of branching and<br />

contact between gametangia (Mullins, 1973; Gow<br />

& Gooday, 1987). A cy<strong>to</strong>plasmic recep<strong>to</strong>r protein<br />

for antheridiol has been detected (Riehl et al.,<br />

1984), and the hormone probably acts like<br />

its equivalents in mammalian cells, by the<br />

recep<strong>to</strong>r hormone complex moving <strong>to</strong> the<br />

nucleus and binding specifically <strong>to</strong> DNA, increasing<br />

transcription rates of certain genes (Elliott,<br />

1994).<br />

There is evidence that the co-ordination of<br />

sexual reproduction by hormonal control is not<br />

confined <strong>to</strong> heterothallic forms of Achlya, but<br />

also takes place in homothallic species. The fact<br />

that it is possible <strong>to</strong> initiate sexual reactions<br />

between homothallic and heterothallic species<br />

of Achlya shows that some of the hormones are<br />

common <strong>to</strong> more than one species, although<br />

Table 5.2. Postulated effects of hormones on sexual reactions in Achlya ambisexualis.<br />

Hormone Produced by Affecting Specific action<br />

A Vegetative hyphae Vegetative hyphae Induces formation of antheridial branches.<br />

B Antheridial branches Vegetative hyphae Initiates formation of oogonial initials.<br />

C Oogonialinitials Antheridial branches (1) Attracts antheridial branches.<br />

(2) Induces thigmotropic response and<br />

delimitation of antheridia.<br />

D Antheridia Oogonialinitials Induces delimitation of oogonium<br />

by formation of basal septum.<br />

After Raper (1939). So far, only hormones A and C (antheridiol) and B (oogoniol) have been shown<br />

<strong>to</strong> exist.


SAPROLEGNIALES<br />

91<br />

there is also evidence of some degree of<br />

specificity of the hormones of different species<br />

(Raper, 1950; Barksdale, 1965).<br />

One further interesting phenomenon which<br />

has been discovered in relation <strong>to</strong> heterothallic<br />

Achlya spp. is relative sexuality. If isolates of<br />

A. bisexualis and A. ambisexualis from separate<br />

sources are paired in all possible combinations, it<br />

is found that certain strains show a capacity <strong>to</strong><br />

react either as male or as female, depending on<br />

the particular partner <strong>to</strong> which they are apposed.<br />

Other strains remain invariably male or invariably<br />

female, and these are referred <strong>to</strong> as true or<br />

strong males or females. The strains can be<br />

arranged in a series with strong males and<br />

strong females at the extremes, and intermediate<br />

strains whose reaction may be either male or<br />

female depending on the strength of their<br />

mating partner. Similar interspecific responses<br />

between strains of A. bisexualis and A. ambisexualis<br />

are also possible. Further, some of the strains<br />

which appear heterothallic at room temperature<br />

are homothallic at lower temperatures.<br />

Barksdale (1960) has postulated that the heterothallic<br />

forms are derived from homothallic ones.<br />

She argued that the most notable difference<br />

between strong males and strong females lies in<br />

their differential antheridiol production and<br />

response. Very little of this substance is found<br />

in male cultures, and these are much more<br />

sensitive in their response <strong>to</strong> the hormone than<br />

female cultures. Another important difference is<br />

in the uptake of antheridiol. Certain strains<br />

appear capable of absorbing it much more<br />

readily than others, and it is the strains with a<br />

high ability <strong>to</strong> absorb antheridiol that produce<br />

antheridial branches during conjugation with<br />

other thalli (Barksdale, 1963). If one assumes that<br />

heterothallic forms have been derived from<br />

homothallic ones, this might have occurred by<br />

mutations leading <strong>to</strong> increased sensitivity <strong>to</strong><br />

antheridiol and hence <strong>to</strong> maleness. Conversely,<br />

mutations leading <strong>to</strong> enhanced extracellular<br />

accumulation of antheridiol should lead <strong>to</strong><br />

increasing femaleness.<br />

Germination of the oospores of A. ambisexualis<br />

results in the formation of a multinucleate germ<br />

tube which develops in<strong>to</strong> a germ sporangium<br />

if transferred <strong>to</strong> water, or in<strong>to</strong> a coenocytic<br />

mycelium in the presence of nutrients. This<br />

mycelium can be induced <strong>to</strong> form zoosporangia<br />

when transferred <strong>to</strong> water. From zoosporangia of<br />

either source, single zoospore cultures can be<br />

obtained which can be mated with the parental<br />

male or female strains. All zoospores or germ<br />

tubes derived from a single oospore gave the<br />

same result with regard <strong>to</strong> their sexual interaction.<br />

This finding suggests that nuclear division<br />

on oospore germination is not meiotic, and is<br />

thus consistent with the idea that the life cycle is<br />

diploid (Mullins & Raper, 1965). Confirmation of<br />

these results, implying meiosis during gamete<br />

differentiation, has also been obtained with<br />

A. ambisexualis (Barksdale, 1966).<br />

5.2.3 Thraus<strong>to</strong>theca, Dictyuchus and<br />

Pythiopsis (Saprolegniaceae)<br />

In Thraus<strong>to</strong>theca clavata the sporangia are broadly<br />

club-shaped, and there is no free-swimming<br />

auxiliary zoospore stage. Encystment occurs<br />

within the sporangia and the auxiliary cysts are<br />

released by irregular rupture of the sporangial<br />

wall (Fig. 5.12a). After release, the angular cysts<br />

germinate <strong>to</strong> release bean-shaped principal zoospores<br />

with laterally attached flagella (Figs.<br />

5.12c,d). After a period of swimming, further<br />

encystment occurs, followed by germination by<br />

a germ tube (Figs. 5.12e,f), or by emergence of a<br />

further principal zoospore. The zoospores are<br />

thus monomorphic and polyplanetic. Sexual<br />

reproduction is homothallic, but formation of<br />

gametangia is stimulated by Achlya sex hormones<br />

(Raper, 1950). Oospores germinate either by a<br />

germ tube or by a germ sporangium (Fig. 5.12g).<br />

In Dictyuchus, there is again no free-swimming<br />

auxiliary zoospore stage. Commonly the entire<br />

zoosporangium is deciduous, and detached zoosporangia<br />

are capable of forming zoospores.<br />

Auxiliary zoospore initials are cleaved out but<br />

encystment occurs within the cylindrical sporangium.<br />

The cysts are tightly packed <strong>to</strong>gether<br />

and release their principal zoospores independently<br />

through separate pores in the sporangial<br />

wall (Fig. 5.13a). When zoospore release is<br />

complete, a network made up of the polygonal<br />

walls of the auxiliary cysts is left behind. After<br />

swimming, the laterally biflagellate zoospores


92 STRAMINIPILA: OOMYCOTA<br />

Fig 5.12 Thraus<strong>to</strong>theca clavata.<br />

(a) Zoosporangium showing formation of<br />

auxiliary cysts within the sporangium.<br />

Theauxiliarycystsarebeingreleasedthrough<br />

breakdown of the sporangial wall.<br />

(b) Auxiliary cyst. (c) Auxiliary cyst<br />

germinating <strong>to</strong> release a principal zoospore,<br />

the first motile stage in this species.<br />

(d) Principal zoospore. (e) Principal cyst.<br />

(f) Principal cyst germinating by means of<br />

a germ tube. (g) Sexual reproduction.<br />

Six-month-old oospore germinating after17 h<br />

in charcoal water.The germ tube is terminated<br />

by a germ sporangium. Bar¼20 mm(a)or<br />

10 mm(b) (g).<br />

encyst (Figs. 5.13b,c). Electron micrographs have<br />

shown that the wall of the secondary cyst of<br />

D. sterile bears a series of long spines looking<br />

somewhat like the fruit of a horse chestnut<br />

(Fig. 5.14; Heath et al., 1970). Following the<br />

formation of the first zoosporangium, a second<br />

may be produced immediately beneath it by the<br />

formation of a septum cutting off a subterminal<br />

segment of the original hypha, or growth may<br />

be renewed laterally <strong>to</strong> the first sporangium<br />

(Fig. 5.13a).<br />

Because there is only one motile stage in<br />

Thraus<strong>to</strong>theca and Dictyuchus (i.e. a zoospore of<br />

the principal type), they are said <strong>to</strong> be monomorphic.<br />

Pythiopsis cymosa (Figs. 5.13e i) is also<br />

monomorphic, but in this species the only<br />

motile stage is of the auxiliary type and principal<br />

zoospores are not formed. After swimming,<br />

the zoospore encysts and then<br />

germinates directly by means of a germ tube<br />

(Figs. 5.13g i).<br />

5.2.4 Aplanetic forms<br />

In certain cultures of Saprolegniaceae the zoosporangia<br />

produce cysts which do not release any<br />

motile stage. Instead, germ tubes are put out<br />

which penetrate the sporangial wall. Forms<br />

without motile spores are said <strong>to</strong> be aplanetic.<br />

The aplanetic condition is occasionally found<br />

in staling cultures of Saprolegnia, Achlya and


SAPROLEGNIALES<br />

93<br />

Fig 5.13 (a d) Dictyuchus sterile. (a) Zoosporangium showing cysts within the sporangium and the release of principal zoospores<br />

through separate pores in the sporangium wall. Note the network of auxiliary cyst walls. (b) Principal zoospores. (c) Principal cyst.<br />

(d) Germination of principal cysts by means of germ tubes. (e i) Pythiopsis cymosa. (e) Zoosporangium. (f,g) Auxiliary zoospores.<br />

(h) Auxiliary cyst. (i) Auxiliary cyst germinating by means of a germ tube. Principal zoospores have not been described.<br />

(a c,e,f) <strong>to</strong> same scale; (g i) <strong>to</strong> same scale.<br />

Dictyuchus. Some species produce sporangia only<br />

rarely and the genus Aplanes has been erected for<br />

these forms. However, in very clean cultural<br />

conditions, all have been shown <strong>to</strong> behave as<br />

Achlya, and they are currently accommodated<br />

within that genus (Dick, 2001a). Two species of<br />

Saprolegniaceae are not known <strong>to</strong> form sporangia<br />

at all. They are common in soil, and have<br />

been placed in a separate genus, Aplanopsis.<br />

Another genus, Geolegnia, forms sporangia containing<br />

thick-walled aplanospores which never<br />

produce a flagellate stage. The final classification<br />

of these small genera of Saprolegniaceae will<br />

have <strong>to</strong> await the results of comparisons of<br />

suitable DNA sequences (see M. A. Spencer et al.,<br />

2002).


94 STRAMINIPILA: OOMYCOTA<br />

Fig 5.14 Surface of a principal cyst of Dictyuchus<br />

sterile.Note the spines covering the surface.Image<br />

kindly provided by M.W. Dick and I.C. Hallett;<br />

reprinted from Hallett and Dick (1986), with<br />

permission from Elsevier.<br />

5.2.5 Aphanomyces (Lep<strong>to</strong>legniaceae)<br />

Aphanomyces is distinguished from Achlya by its<br />

thin, delicate hyphae and its narrow sporangia<br />

containing a single row of spores. Based on these<br />

morphological differences and DNA sequence<br />

analyses, the genus Aphanomyces has been<br />

removed from the Saprolegniaceae and classified<br />

in the family Lep<strong>to</strong>legniaceae, still within the<br />

Saprolegniales (Dick et al., 1999; Hudspeth et al.,<br />

2000; Dick, 2001a).<br />

Asexual reproduction in Aphanomyces is variable.<br />

In A. euteiches, flagella do not develop on the<br />

first-formed spores. Pro<strong>to</strong>plasts are cleaved out,<br />

move <strong>to</strong> the mouth of the sporangium, and<br />

encyst. Principal zoospores develop from the<br />

cysts and are the first true motile stage. Aphanomyces<br />

euteiches is thus monomorphic. In A. patersonii,<br />

the motility of the first-formed zoospore is<br />

controlled by variation in temperature. Below<br />

20°C, encystment of the auxiliary zoospores at the<br />

mouth of the sporangium occurs in a manner<br />

typical of the genus, but above this temperature<br />

the auxiliary zoospores swim away and encyst<br />

some distance away from the zoosporangium.<br />

The genus Aphanomyces has been monographed<br />

by Scott (1961). It has gained no<strong>to</strong>riety<br />

particularly because A. astaci is the cause of<br />

the plague of European crayfish. Having been<br />

introduced probably in the 1860s from America,<br />

where the local crayfish populations are fairly<br />

resistant <strong>to</strong> A. astaci infections, the fungus has<br />

now spread across Europe, severely damaging<br />

commercial production of the highly susceptible<br />

European crayfish, Astacus fluviatilis<br />

(Alderman & Polglase, 1986; Cerenius et al.,<br />

1988; Alderman et al., 1990). Although it would<br />

be possible <strong>to</strong> introduce resistant s<strong>to</strong>ck of<br />

American crayfish in<strong>to</strong> European river systems<br />

affected by the disease, resistant crayfish still<br />

harbour the pathogen, thereby making it impossible<br />

<strong>to</strong> res<strong>to</strong>re the native crayfish populations<br />

in the future (Dick, 2001a). The difference in<br />

resistance between North American and European<br />

crayfish lies in the melanization reaction<br />

which arrests hyphal growth from encysted<br />

zoospores (Nyhlén & Unestam, 1980; Cerenius<br />

et al., 1988). In European crayfish, melanization<br />

occurs <strong>to</strong>o slowly <strong>to</strong> prevent the spread of the<br />

fungus in<strong>to</strong> the haemocoel which causes rapid<br />

death. Aphanomyces astaci can also cause epizootic<br />

ulcerative disease in fish, the symp<strong>to</strong>ms often<br />

being very similar <strong>to</strong> those caused by Saprolegnia<br />

(Lilley & Roberts, 1997).<br />

Aphanomyces euteiches is a significant pathogen<br />

of roots of peas and other terrestrial plants<br />

(Papavizas & Ayers, 1974; Persson et al., 1997).<br />

Recently, methods have been developed <strong>to</strong> quantify<br />

the prevalence of the pathogen in infected<br />

plants by measuring the levels of specific fatty<br />

acids which are produced by A. euteiches but not by<br />

plants or pathogens belonging <strong>to</strong> the Eumycota<br />

(Larsen et al., 2000). Other species of Aphanomyces


PYTHIALES<br />

95<br />

are keratinophilic, occurring in the soil or in<br />

water on insect remains (Dick, 1970; Seymour &<br />

Johnson, 1973).<br />

5.3 Pythiales<br />

The order Pythiales includes two families, the<br />

Pythiaceae and Pythioge<strong>to</strong>naceae (Dick, 2001a;<br />

Kirk et al., 2001). The Pythioge<strong>to</strong>naceae are a small<br />

group of aquatic saprotrophs presently comprising<br />

one genus and six species. They occur in<br />

anoxic sediments at the bot<strong>to</strong>m of freshwater<br />

lakes and are facultatively anaerobic as well as<br />

obligately fermentative, i.e. they break down<br />

sugars incompletely <strong>to</strong> give organic acids irrespective<br />

of the presence or absence of oxygen<br />

(Emerson & Natvig, 1981; Natvig & Gleason, 1983).<br />

Another member of the Pythioge<strong>to</strong>naceae,<br />

Pythioge<strong>to</strong>n zeae, causes root and stalk rot in<br />

maize (Jee et al., 2000). The Pythioge<strong>to</strong>naceae are<br />

clearly related <strong>to</strong> the Pythiaceae by DNA sequence<br />

homology (Voglmayr et al., 1999).<br />

Only the Pythiaceae will be considered<br />

further in this book. This is a large family of<br />

over 200 species in approximately 10 genera, of<br />

which 2 are of outstanding significance: Pythium<br />

and Phy<strong>to</strong>phthora. Phy<strong>to</strong>phthora species are primarily<br />

pathogenic <strong>to</strong> plants from which they can be<br />

isolated and grown in pure culture. The genus<br />

Pythium is best known for its saprotrophic soilinhabiting<br />

members, many of which are opportunistic<br />

pathogens especially in young plants.<br />

There are also obligately pathogenic Pythium<br />

spp. Generally, Pythium spp. parasitize a wider<br />

diversity of hosts than Phy<strong>to</strong>phthora, including<br />

mammals, fungi and algae.<br />

5.3.1 Life cycle of Pythiaceae<br />

The life cycle of Phy<strong>to</strong>phthora infestans is summarized<br />

in Fig. 5.19. Asexual reproduction in Pythium<br />

and Phy<strong>to</strong>phthora is by means of sporangia<br />

which vary in shape from swollen hyphae or<br />

globose structures (Pythium) <strong>to</strong> lemon-shaped<br />

(Phy<strong>to</strong>phthora). Sporangia are borne on more or<br />

less undifferentiated hyphae. In most cases,<br />

sporangia germinate <strong>to</strong> produce zoospores<br />

which are of the principal (kidney-shaped) type.<br />

In many Pythium spp., the final stages of zoospore<br />

differentiation take place outside the sporangium<br />

in a walled vesicle, followed by breakdown<br />

of the soft wall and release of the zoospores. In<br />

Phy<strong>to</strong>phthora, in contrast, zoospores differentiate<br />

within the sporangium and are released directly<br />

or via a very short-lived vesicle which is surrounded<br />

only by a membrane. About 20% of<br />

the <strong>to</strong>tal respira<strong>to</strong>ry activity within a released<br />

zoospore is used up <strong>to</strong> fuel propulsion (Hölker<br />

et al., 1993). The forward-directed straminipilous<br />

flagellum generates about 10 times more thrust<br />

than the posterior whiplash flagellum which acts<br />

mainly as a rudder (Erwin & Ribeiro, 1996).<br />

Zoospores can swim for several hours before they<br />

encyst. The process of encystment has been<br />

examined in great detail for Phy<strong>to</strong>phthora (see<br />

p. 102). Cysts usually germinate by means of<br />

a germ tube, only rarely producing a further<br />

zoospore stage. In many species, sporangia can<br />

germinate either indirectly by releasing zoospores<br />

or directly by means of a germ tube,<br />

depending on environmental conditions and<br />

age of the sporangium.<br />

Sexual reproduction is oogamous. Each oogonium<br />

contains a single oosphere (except for<br />

Pythium multisporum in which there are several).<br />

The antheridial and oogonial initials are<br />

commonly multinucleate at their inception and<br />

further nuclear divisions may occur during<br />

development. Meiosis eventually takes place in<br />

the gametangia so that karyogamy occurs<br />

between haploid antheridial and oogonial<br />

nuclei. In many forms, there is only one functional<br />

male and female nucleus, but in others<br />

multiple fusions occur. Oospores germinate<br />

either by producing a single germ sporangium,<br />

or by sending out vegetative hyphae.<br />

Most members of the Pythiaceae are homothallic,<br />

although heterothallism and relative<br />

sexuality have been reported, e.g. for<br />

Phy<strong>to</strong>phthora infestans (Fig. 5.19) and Pythium<br />

sylvaticum. Heterothallic species are thought <strong>to</strong><br />

be derived from homothallic ones (Kroon et al.,<br />

2004). The situation of mating in heterothallic<br />

strains is rather complex and still only incompletely<br />

unders<strong>to</strong>od. A system of two mating types<br />

(A1 and A2) seems <strong>to</strong> be superimposed on a<br />

hormonal control mechanism of mating akin <strong>to</strong>


96 STRAMINIPILA: OOMYCOTA<br />

that described for Achlya (p. 86). When two strains<br />

of Pythium or Phy<strong>to</strong>phthora were separated by a<br />

membrane preventing hyphal contact but<br />

permitting the exchange of diffusible metabolites,<br />

oospores were formed by either or both<br />

strains (Ko, 1980; Gall & Elliott, 1985). Because<br />

the mycelia were separated by a membrane,<br />

oospores formed by selfing, whereas in direct<br />

contact they may form by hybridization<br />

(Shat<strong>to</strong>ck et al., 1986a,b). Oospore formation can<br />

also be induced by non-specific stimuli, such as<br />

volatile metabolites of the unrelated fungus<br />

Trichoderma stimulating reproduction in A2 but<br />

not A1 strains of Phy<strong>to</strong>phthora palmivora (Brasier,<br />

1975a). This ‘Trichoderma effect’ may well have<br />

ecological implications, since Trichoderma spp.<br />

are very common, especially in soil. Oospore<br />

formation may be a defence reaction against<br />

antibiotics commonly produced by Trichoderma,<br />

and the ‘Trichoderma effect’ may actually enhance<br />

the survival of Phy<strong>to</strong>phthora spp. in soil, since it<br />

stimulates production of the long-lived oospore<br />

stage even in the absence of a compatible mating<br />

type (Brasier, 1975b). It is not known why<br />

Trichoderma spp. do not stimulate oosporogenesis<br />

in A1 strains.<br />

Like Achlya, the Pythiaceae display relative<br />

sexuality, i.e. a strain can act as male in one<br />

pairing but as female in another. To complicate<br />

matters further, a given strain of Phy<strong>to</strong>phthora<br />

parasitica can switch its mating type from<br />

predominantly male <strong>to</strong> predominantly female<br />

or vice versa, e.g. upon fungicide treatment<br />

(Ko et al., 1986). Clearly, despite substantial<br />

research efforts over many years the genetic<br />

basis of sexual reproduction in the Pythiaceae<br />

still poses numerous unresolved questions!<br />

By analogy with the hormones oogoniol and<br />

antheridiol of Achlya, a male strain needs <strong>to</strong> be<br />

induced <strong>to</strong> produce the oogonium-inducing<br />

hormone whereas female strains constitutively<br />

produce the antheridium-inducing hormone<br />

(Elliott, 1994). The ability of homothallic species<br />

<strong>to</strong> stimulate sexual reproduction in heterothallic<br />

species (Ko, 1980) indicates that these hormones<br />

may also fulfil a morphogenetic role in homothallic<br />

sexual reproduction. However, nothing<br />

seems <strong>to</strong> be known as yet about the chemical<br />

nature of these hormones.<br />

Sterols are neither synthesized nor strictly<br />

required by vegetatively growing Pythium or<br />

Phy<strong>to</strong>phthora spp. (Nes et al., 1979). None the less,<br />

they are required for the formation of sexual<br />

reproductive organs (Elliott, 1994). It seems,<br />

therefore, that sterols especially si<strong>to</strong>sterol<br />

and stigmasterol which are normally taken up<br />

from the host plant are converted in<strong>to</strong> as<br />

yet unidentified steroid hormones which initiate<br />

sexual morphogenetic events downstream of<br />

the action of the diffusible Achlya-like hormones<br />

(Elliott, 1994). An alternative hypothesis is that<br />

sterols interact with an as yet unknown membrane<br />

protein <strong>to</strong> transmit the hormonal signal<br />

and trigger the signalling cascade leading <strong>to</strong><br />

sexual morphogenesis (Nes & Stafford, 1984).<br />

In Lagenidium giganteum, a member of the<br />

Pythiaceae parasitizing mosqui<strong>to</strong> larvae (Cuda<br />

et al., 1997), this cascade seems <strong>to</strong> be carried by<br />

Ca 2þ and calmodulin (Kerwin & Washino, 1986).<br />

5.3.2 Pythium<br />

Species of Pythium grow in water and soil as<br />

saprotrophs, but under suitable conditions, e.g.<br />

where seedlings are grown crowded <strong>to</strong>gether in<br />

poorly drained soil, they can become parasitic,<br />

causing diseases such as pre-emergence killing,<br />

damping off and foot rot. Damping off of cress<br />

(Lepidium sativum) can be demonstrated by sowing<br />

seeds densely on heavy garden soil or garden<br />

compost which is kept liberally watered. Within<br />

5 7 days some of the seedlings may show brown<br />

zones at the base of the hypocotyl, and the<br />

hypocotyl and cotyledons become water-soaked<br />

and flaccid. In this condition the seedling<br />

collapses. A collapsed seedling coming in<strong>to</strong><br />

contact with other seedlings will spread the<br />

disease (Plate 2b). The host cells separate from<br />

each other easily due <strong>to</strong> the breakdown of the<br />

middle lamella, probably brought about by<br />

pectic and possibly cellulolytic enzymes secreted<br />

by the fungus. The enzymes diffuse from their<br />

points of secretion at the hyphal tips, so that<br />

softening of the host tissue actually occurs ahead<br />

of the growing mycelium. Pure culture studies<br />

suggest that species of Pythium may also secrete<br />

heat-stable substances which are <strong>to</strong>xic <strong>to</strong> plants.<br />

Within the host the mycelium is coarse and


PYTHIALES<br />

97<br />

Fig 5.15 Pythium mycelium in the rotting tissue of a cress<br />

seedling hypocotyl. Note the spherical sporangium initial and<br />

the absence of haus<strong>to</strong>ria.<br />

coenocytic, with typically granular cy<strong>to</strong>plasmic<br />

contents (Fig. 5.15). At first there are no septa,<br />

but later cross walls may cut off empty portions<br />

of hyphae. Thick-walled chlamydospores may<br />

also be formed. There are no haus<strong>to</strong>ria.<br />

Several species are known <strong>to</strong> cause damping<br />

off, e.g. P. debaryanum and, perhaps more<br />

frequently, P. ultimum. Pythium aphanidermatum<br />

is associated with stem rot and damping off of<br />

cucumber, and the fungus may also cause rotting<br />

of mature cucumbers. Pythium mamillatum<br />

causes damping off of mustard and beet seedlings<br />

and is also associated with root rot in Viola.<br />

Many Pythium spp. have a very wide host range;<br />

e.g. P. ultimum parasitizes over 150 plant species<br />

belonging <strong>to</strong> many different families (Middle<strong>to</strong>n,<br />

1943; Hendrix & Campbell, 1973). Far from<br />

parasitizing only plant roots, several soil-borne<br />

species, e.g. P. oligandrum, P. acanthicum and<br />

P. nunn, are capable of attacking hyphae of<br />

filamen<strong>to</strong>us fungi, including plant-pathogenic<br />

species and even other Pythium spp. (Foley &<br />

Deacon, 1986b; Deacon et al., 1990). Attack may<br />

be mediated by the secretion of wall-degrading<br />

b-1,3-glucanase, chitinase and cellulase, or by<br />

inducing the host <strong>to</strong> undergo au<strong>to</strong>lysis (Elad<br />

et al., 1985; Laing & Deacon, 1991; Fang & Tsao,<br />

1995). In contrast <strong>to</strong> plant-pathogenic Pythium<br />

spp., the mycoparasitic species require thiamine<br />

for growth and are unable <strong>to</strong> utilize inorganic<br />

nitrogen sources. These deficiencies may<br />

explain their mycoparasitic habit (Foley &<br />

Deacon, 1986a). Other species of Pythium parasitize<br />

freshwater and marine algae (Kerwin et al.,<br />

1992).<br />

The taxonomy of Pythium is somewhat<br />

confused at present due <strong>to</strong> the existence of<br />

numerous synonyms. Including a few varieties,<br />

Dick (2001a) listed 129 names in current use.<br />

Since the morphological characteristics traditionally<br />

used for diagnosis can be variable, the<br />

delimitation of species and their assignment <strong>to</strong><br />

the genus Pythium will have <strong>to</strong> await the results<br />

of detailed molecular phylogenetic analyses<br />

which are in progress (Matsumo<strong>to</strong> et al., 1999;<br />

Lévesque & de Cock, 2004). Keys and descriptions<br />

have been published by Waterhouse (1967,<br />

1968), van der Plaats-Niterink (1981) and Dick<br />

(1990b).<br />

Asexual reproduction<br />

The mycelium within the host tissue or in<br />

culture usually produces sporangia, but their<br />

form varies. In some species, e.g. P. gracile,


98 STRAMINIPILA: OOMYCOTA<br />

Fig 5.16 Sporangia and zoospores of Pythium.(a)Pythium debaryanum. Spherical sporangium with short tube and a vesicle<br />

containing zoospores. (b k) Pythiumaphanidermatum. (b) Lobed sporangium showing a long tube and the vesicle, which is beginning<br />

<strong>to</strong> expand. (c g) Further stages in the enlargement of the vesicle, and differentiation of zoospores.Note the transfer of pro<strong>to</strong>plasm<br />

from the sporangium <strong>to</strong> the vesicle in (c).The stages illustrated in (b g) <strong>to</strong>ok place in 25 min. (h) Enlarged vesicle showing the<br />

zoospores. Flagella are also visible. (i) Zoospores. (j) Encystment of zoospore showing a shed flagellum. (k) Germination of<br />

azoosporecyst.(b g) <strong>to</strong> same scale; (a) and (h k) <strong>to</strong> same scale.<br />

the sporangia are filamen<strong>to</strong>us and are scarcely<br />

distinguishable from vegetative hyphae.<br />

In P. aphanidermatum, the sporangia are formed<br />

from inflated lobed hyphae (Fig. 5.16b). In many<br />

species, however, e.g. P. debaryanum, the sporangia<br />

are globose (Fig. 5.16a). A terminal or<br />

intercalary portion of a hypha enlarges and<br />

assumes a spherical shape, then becomes cut<br />

off from the mycelium by a cross wall. The<br />

sporangia contain numerous nuclei. Cleavage of<br />

the cy<strong>to</strong>plasm <strong>to</strong> form zoospores begins in the<br />

sporangium, but is completed within a thinwalled<br />

vesicle which is extruded from the<br />

sporangium. This is a homohylic vesicle because<br />

its glucan wall is continuous with one layer of<br />

the sporangial wall (Dick, 2001a). Within the<br />

sporangium, cleavage vesicles begin <strong>to</strong> coalesce<br />

<strong>to</strong> separate the cy<strong>to</strong>plasm in<strong>to</strong> uninucleate<br />

portions; membrane-bound packets of TTHs<br />

are already present within the cy<strong>to</strong>plasm of the<br />

sporangium. In P. middle<strong>to</strong>nii (Fig. 5.17), the<br />

fascinating process of differentiation from amorphous<br />

cy<strong>to</strong>plasm <strong>to</strong> motile zoospores takes about<br />

30 45 min (Webster, 2006a) and is readily<br />

demonstrated in the labora<strong>to</strong>ry (Weber et al.,<br />

1999). The sporangium is extended in<strong>to</strong> an apical<br />

papilla capped by a mass of fibrillar material<br />

which is lamellate in ultrastructure (Lunney &<br />

Bland, 1976). Shortly before sporangial discharge,<br />

there is an accumulation of cleavage vesicles<br />

behind the apical cap and at the periphery of<br />

the cy<strong>to</strong>plasm close <strong>to</strong> the sporangium wall. The<br />

cleavage vesicles around the sporangial cy<strong>to</strong>plasm<br />

discharge their contents <strong>to</strong> form a loose,<br />

fibrous interface between the cy<strong>to</strong>plasm and the<br />

sporangial wall.<br />

Discharge of the sporangium occurs by the<br />

formation of a thin-walled vesicle at the tip of<br />

the papilla from the fibrillar material of the<br />

apical cap, and the partially differentiated<br />

zoospore mass is extruded in<strong>to</strong> it. The movement<br />

of the cy<strong>to</strong>plasm from the sporangium in<strong>to</strong> the


PYTHIALES<br />

99<br />

Fig 5.17 Pythium middle<strong>to</strong>nii. Stages in zoospore discharge. (a) Sporangium shortly before discharge. Note the thickened tip of<br />

the papilla which consists of a cap of cell wall material. (b) Inflation of the vesicle begins. (c,d) Pro<strong>to</strong>plasm is retreating from the<br />

sporangium.Note the shrinkage in sporangium diameter as compared with (a). (e) Zoospores have differentiated within the vesicle,<br />

with flagella visible between the vesicle wall and the zoospores. (f) Zoospores escape following the rupture of the vesicle wall.<br />

The whole process of discharge takes about 20 min.<br />

vesicle is probably the result of several forces<br />

including the elastic contraction of the sporangium<br />

wall and possibly surface energy (Webster<br />

& Dennis, 1967). Lunney and Bland (1976) have<br />

also suggested that the fibrillar material<br />

extruded from the cleavage vesicles at the<br />

zoosporangium periphery may imbibe water,<br />

resulting in a build up of turgor pressure. The<br />

vesicle enlarges as cy<strong>to</strong>plasm from the sporangium<br />

is transferred <strong>to</strong> it, and during the next few<br />

minutes the cy<strong>to</strong>plasm cleaves in<strong>to</strong> 8 20 uninucleate<br />

zoospores which jostle about inside the<br />

sporangium, causing the thin vesicle wall <strong>to</strong><br />

bulge irregularly (Fig. 5.17). Finally, about 20 min<br />

after the inflation of the vesicle, its wall breaks<br />

down and the zoospores swim away. Internal<br />

sporangial proliferation, i.e. the formation of a<br />

new sporangium inside an old discharged one,<br />

occurs in certain species, e.g. P. middle<strong>to</strong>nii and<br />

P. undulatum.


100 STRAMINIPILA: OOMYCOTA<br />

In some forms, e.g. P. ultimum var. ultimum,<br />

sporangia do not release zoospores but germinate<br />

directly by producing a germ tube.<br />

Sporangia of P. ultimum var. ultimum may survive<br />

in soil, whether moist or air-dry, for several<br />

months, and are stimulated <strong>to</strong> germinate within<br />

a few hours by sugar-containing exudates from<br />

seed coats. The germ tubes grow very rapidly so<br />

that a host in the vicinity may be penetrated<br />

within 24 h (Stanghellini & Hancock, 1971). The<br />

oospore of P. ultimum var. ultimum can germinate<br />

either by means of a germ tube or by forming<br />

a zoosporangium which releases zoospores<br />

(Figs. 5.18d,e).<br />

The zoospore<br />

Zoospores of Pythium spp. are always of the<br />

principal type. They can swim for several hours<br />

in a readily recognizable manner of helical<br />

forward movement. Donaldson and Deacon<br />

(1993) have provided evidence that the zoospore<br />

swimming pattern is regulated by Ca 2þ and<br />

calmodulin; manipulations of Ca 2þ concentrations<br />

cause aberrations such as circular, straight,<br />

spirally skidding or irregular movement. Zoospores<br />

of Pythium are attracted <strong>to</strong>wards host<br />

surfaces, usually roots. The Ca 2þ /calmodulin<br />

system may be the means by which the sensing<br />

of attractants is translated in<strong>to</strong> directed movement.<br />

It is this directed movement (taxis), i.e. the<br />

ability <strong>to</strong> aim precisely at a suitable encystment<br />

site, rather than the ability <strong>to</strong> move per se,<br />

which represents the main benefit of zoospores<br />

<strong>to</strong> their producer (Deacon & Donaldson, 1993).<br />

Chemotaxis <strong>to</strong> root exudates is often nonhost-specific,<br />

being mediated by amino acids<br />

and other common metabolites ( Jones et al.,<br />

1991). Other tactic movements also occur, such<br />

as pho<strong>to</strong>taxis, electrotaxis or negative geotaxis<br />

(Dick, 2001a). In general, zoospores of Pythium<br />

spp. accumulate around the root cap, root<br />

elongation zone or sites of injury.<br />

Once the zoospore has alighted on a suitable<br />

surface, it encysts by shedding rather than<br />

withdrawing its flagella, and secreting a wall<br />

from pre-formed material. Much valuable ultrastructural<br />

work has been carried out on the<br />

encystment process of Phy<strong>to</strong>phthora and is<br />

discussed on pp. 102 111. The cyst of Pythium<br />

spp. can germinate almost immediately, usually<br />

by emitting a germ tube which can directly<br />

penetrate the relatively soft root tissue. In P.<br />

marinum, which is parasitic on marine red algae,<br />

the germ tube forms a specialized infection<br />

structure termed an appressorium (Kerwin<br />

et al., 1992); this is also commonly formed by<br />

leaf-infecting Phy<strong>to</strong>phthora spp. The entire process<br />

from zoospore encystment <strong>to</strong> successful penetration<br />

is called homing sequence and may take<br />

place in as little as 30 min (Deacon & Donaldson,<br />

1993). If a zoospore encysts on a non-host surface,<br />

the cyst may germinate by producing a further<br />

principal zoospore.<br />

Sexual reproduction<br />

Most species of Pythium are homothallic, i.e.<br />

oogonia and antheridia are readily formed in<br />

cultures derived from single zoospores. However,<br />

Fig 5.18 Oogonia and oospores of<br />

Pythium.(a)Pythium debaryanum.<br />

Note that there are several antheridia.<br />

(b) Pythium mamillatum.Oogonium<br />

showing spiny outgrowths of oogonial<br />

wall. (c) Pythium ultimum.<br />

(d, e) Germination of oospores of<br />

P. ultimum (after Drechsler,1960).


PYTHIALES<br />

101<br />

Fig 5.19 Life cycle of Phy<strong>to</strong>phthora infestans.This fungus is heterothallic, and the asexual part of the life cycle (left of diagram) is<br />

shown only for one mating type (A1). Nuclei in vegetative states are diploid.When two compatible mycelia meet, multinucleate<br />

oogonia and antheridia are differentiated, and one meiotic event in each results in the transfer of one haploid nucleus from the<br />

gametangium <strong>to</strong> the oogonium. Karyogamy is delayed until shortly before oospore germination.Open and closed circles represent<br />

haploid nuclei of opposite mating type; diploid nuclei are larger and half-filled. Key events in the life cycle are meiosis (M),<br />

plasmogamy (P) and karyogamy (K).<br />

some heterothallic species are known, e.g.<br />

P. sylvaticum, P. heterothallicum and P. splendens.<br />

In these cases, mating is a complicated affair<br />

under hormonal control, and with relative<br />

sexuality (see p. 95).<br />

Oogonia arise as terminal or intercalary<br />

spherical swellings which become cut off from<br />

the adjacent mycelium by cross-wall formation.<br />

In some species, e.g. P. mamillatum, the oogonial<br />

wall is folded in<strong>to</strong> long projections (Fig. 5.18b).<br />

The antheridia arise as club-shaped swollen<br />

hyphal tips, often as branches of the oogonial<br />

stalk (monoclinous) or sometimes from separate<br />

hyphae (diclinous). In some species, e.g.<br />

P. ultimum, there is typically only a single antheridium<br />

<strong>to</strong> each oogonium, whilst in others, e.g.<br />

P. debaryanum, there may be several (Fig. 5.18a).<br />

The young oogonium is multinucleate and<br />

the cy<strong>to</strong>plasm within it differentiates in<strong>to</strong> a<br />

multinucleate central mass, the ooplasm from<br />

which the oosphere develops, and a peripheral<br />

mass, the periplasm, also containing several<br />

nuclei. The periplasm does not contribute <strong>to</strong><br />

the formation of the oosphere.<br />

As soon as the gametangia have become<br />

delimited by the basal septum, mi<strong>to</strong>tic divisions<br />

cease. Nuclei may be aborted at this stage, and<br />

in oogonia of P. debaryanum 1 8 nuclei undergo<br />

meiosis (Sansome, 1963). Meiotic divisions are<br />

synchronous in the antheridium and the oogonium,<br />

although no pro<strong>to</strong>plasmic continuities<br />

exist at this stage (Dick, 1995). In the antheridium<br />

of P. debaryanum and P. ultimum, all nuclei<br />

but one degenerate prior <strong>to</strong> meiosis, so that four


102 STRAMINIPILA: OOMYCOTA<br />

haploid nuclei are present in each antheridium<br />

just prior <strong>to</strong> plasmogamy (Sansome, 1963; Win-<br />

Tin & Dick, 1975). The antheridium then attaches<br />

itself <strong>to</strong> the oogonial wall and penetrates it by<br />

means of a fertilization tube. Following penetration,<br />

only three nuclei were counted in the<br />

antheridium, suggesting that one had entered<br />

the oogonium. Later still, empty antheridia were<br />

found, and it is presumed that the three<br />

remaining nuclei enter the oogonium and join<br />

the oogonial nuclei degenerating in the periplasm.<br />

Fusion between a single antheridial and<br />

oosphere nucleus has been described. The fertilized<br />

oosphere secretes a double wall, and the<br />

ooplast appears in the pro<strong>to</strong>plasm. Material<br />

derived from the periplasm may also be deposited<br />

on the outside of the developing oospore.<br />

Such oospores may need a period of rest (afterripening)<br />

of several weeks before they are<br />

capable of germinating. Germination may be<br />

by means of a germ tube, or by the formation of<br />

a vesicle in which zoospores are differentiated<br />

(Figs. 5.18d,e), or in some forms the germinating<br />

oospore produces a short germ tube terminating<br />

in a sporangium.<br />

Ecological considerations<br />

Pythium spp. can live saprotrophically and may<br />

survive in air-dry soil for several years. They<br />

are more common in cultivated than in natural<br />

soils (Foley & Deacon, 1985), and appear <strong>to</strong> be<br />

in<strong>to</strong>lerant of highly acidic soils. As saprotrophs,<br />

species of Pythium are important primary colonizers,<br />

probably gaining initial advantage by<br />

virtue of their rapid growth rate. They do not,<br />

however, compete well with other fungi which<br />

have already colonized a substrate, and they<br />

appear <strong>to</strong> be rather in<strong>to</strong>lerant of antibiotics.<br />

The control of diseases caused by Pythium is<br />

obviously rendered difficult by its ability <strong>to</strong><br />

survive saprotrophically and as oospores in soil.<br />

Its wide host range means that it is not possible<br />

<strong>to</strong> control diseases by means of crop rotation.<br />

The effects of disease can be reduced by improving<br />

drainage and avoiding overcrowding of<br />

seedlings. Pythium infections are particularly<br />

severe in greenhouses and nurseries, where<br />

some measure of control can be achieved by<br />

partial steam sterilization of soil. Recolonization<br />

of the treated soil by Pythium is slow. The use of<br />

certain types of compost instead of peat in<br />

nurseries can provide good control (Craft &<br />

Nelson, 1996; Zhang et al., 1996). The fungicide<br />

metalaxyl (see Fig. 5.27) also gives good control<br />

of seedling blight.<br />

Pythium insidiosum<br />

This species is associated with algae in stagnant<br />

freshwater in tropical and subtropical regions.<br />

When horses or cattle come in<strong>to</strong> contact with<br />

P. insidiosum-contaminated water, zoospores are<br />

attracted <strong>to</strong> wounds and can infect them,<br />

causing severe open lesions of skin and subcutaneous<br />

tissues known as pythiosis insidiosi (Meireles<br />

et al., 1993; Mendoza et al., 1993). If contaminated<br />

water is consumed, gastrointestinal or systemic<br />

infections may also arise. In addition <strong>to</strong> grazing<br />

animals, infections in dogs and humans have<br />

been reported. Pythium insidiosum is keratinophilic<br />

and survives well at 37°C. Infections can be<br />

treated successfully by immunotherapy in which<br />

horses are injected with killed fungal material,<br />

the immune response leading <strong>to</strong> healing of<br />

infections (Mendoza et al., 1992). Pythium insidiosum<br />

used <strong>to</strong> be known under different names,<br />

but its taxonomy has been clarified by de Cock<br />

et al. (1987).<br />

5.3.3 Phy<strong>to</strong>phthora<br />

The name Phy<strong>to</strong>phthora (Gr.: ‘plant destroyer’) is<br />

apt, most species being highly destructive plant<br />

pathogens. The best known is P. infestans, cause of<br />

late blight of pota<strong>to</strong>es (Plate 2e). This fungus is<br />

confined <strong>to</strong> solanaceous hosts (especially <strong>to</strong>ma<strong>to</strong><br />

and pota<strong>to</strong>), but others have a much wider host<br />

range. For example, P. cac<strong>to</strong>rum has been recorded<br />

from over 200 species belonging <strong>to</strong> 60 families of<br />

flowering plants, causing a variety of diseases<br />

such as damping off or rots of roots, fruits<br />

and shoots (Erwin & Ribeiro, 1996). Phy<strong>to</strong>phthora<br />

cinnamomi has the widest host range of<br />

all species, being capable of infecting over 1000<br />

plants and causing serious diseases especially on<br />

woody hosts, including conifers and Eucalyptus<br />

(Zentmyer, 1980). Several other Phy<strong>to</strong>phthora spp.<br />

and related Pythium spp. can also cause diebacks<br />

and sudden-death symp<strong>to</strong>ms of trees, with


PYTHIALES<br />

103<br />

roots severely rotted by the time above-ground<br />

symp<strong>to</strong>ms become apparent (Plate 2c,d). Other<br />

important pathogens are P. erythroseptica associated<br />

with pink rot of pota<strong>to</strong> tubers (Plate 2f),<br />

P. fragariae causing red core of strawberries,<br />

and P. palmivora causing pod rot and canker of<br />

cocoa. The genus is cosmopolitan, although there<br />

are differences in the geographic distribution<br />

of individual species; for instance, P. cac<strong>to</strong>rum,<br />

P. nicotianae, P. cinnamomi and P. drechsleri occur<br />

worldwide whereas P. fragariae and P. erythroseptica<br />

are found predominantly in Northern<br />

Europe and North America (Erwin & Ribeiro,<br />

1996). Many Phy<strong>to</strong>phthora spp. are spreading<br />

actively at present, e.g. P. infestans which has<br />

been spread worldwide by human activity (Fry &<br />

Goodwin, 1997) or P. ramorum, a serious pathogen<br />

of oak trees and other woody plants (Henricot &<br />

Prior, 2004). To make matters worse, different<br />

Phy<strong>to</strong>phthora species may hybridize in nature,<br />

producing strains with new host spectra. An<br />

example is the recent outbreak of wilt of Alnus<br />

glutinosa in Europe caused by P. alni, a tetraploid<br />

hybrid of species resembling P. cambivora and<br />

P. fragariae (Brasier et al., 2004).<br />

In accordance with the great importance of<br />

the genus Phy<strong>to</strong>phthora in mycology and plant<br />

pathology, a vast amount of literature has been<br />

published, and some of it has been summarized<br />

by Erwin & Ribeiro (1996) and Dick (2001a).<br />

Several books on the genus have appeared,<br />

including those edited by Erwin et al. (1983),<br />

Ingram and Williams (1991) and Lucas et al.<br />

(1991), and the masterly compendium by Erwin<br />

and Ribeiro (1996). Keys <strong>to</strong> the genus have been<br />

produced by Waterhouse (1963, 1970) and<br />

Stamps et al. (1990). Including formae speciales,<br />

Dick (2001a) listed 84 names in current use.<br />

Phy<strong>to</strong>phthora is closely related <strong>to</strong> Pythium and<br />

there are transitional species which may need <strong>to</strong><br />

be re-assigned as more DNA sequences and other<br />

data become available (Panabières et al., 1997).<br />

In general, the two genera can be distinguished<br />

morphologically in that the sporangia of<br />

Phy<strong>to</strong>phthora spp. are typically pear- or lemonshaped<br />

with an apical papilla (Fig. 5.20b), and<br />

ecologically by the predominantly saprotrophic<br />

existence of Pythium and the predominantly<br />

parasitic mode-of-life of Phy<strong>to</strong>phthora. Probably<br />

all Phy<strong>to</strong>phthora spp. are pathogenic on plants in<br />

some form, and they differ merely in the extent<br />

<strong>to</strong> which they have a free-living saprotrophic<br />

phase. All may survive in the soil at least in<br />

the form of oospores, or in infected host tissue.<br />

However, in contrast <strong>to</strong> the downy mildews<br />

(Peronosporales; Section 5.4), almost all pathogenic<br />

forms can be isolated from their hosts and<br />

can be grown in pure culture. Selective media,<br />

often incorporating antibiotics or fungicides<br />

such as pimaricin or benomyl, have been devised<br />

for the isolation of Phy<strong>to</strong>phthora (Tsao, 1983;<br />

Erwin & Ribeiro, 1996).<br />

Vegetative growth<br />

Most species form an aseptate mycelium producing<br />

branches at right angles, often constricted<br />

at their point of origin. Septa may be present in<br />

older cultures. Within the host, the mycelium is<br />

intercellular, but haus<strong>to</strong>ria may be formed.<br />

These are specialized hyphal branches which<br />

penetrate the wall of the host cell and invaginate<br />

its plasmalemma, thereby establishing a point of<br />

contact between pathogen and host membranes.<br />

Haus<strong>to</strong>ria are typical of biotrophic pathogens<br />

such as the Peronosporales (see Fig. 5.29) but may<br />

also be formed during initial biotrophic phases<br />

of infections which subsequently turn necrotrophic.<br />

In P. infestans within pota<strong>to</strong> tubers, the<br />

haus<strong>to</strong>ria appear as finger-like protuberances<br />

(Fig. 5.20c). Electron micrographs of infected<br />

pota<strong>to</strong> leaves show that the haus<strong>to</strong>ria are not<br />

surrounded by host cell wall material, but by an<br />

encapsulation called the extrahaus<strong>to</strong>rial matrix<br />

which is probably of fungal origin. This is<br />

delimited on the outside by the host plasma<br />

membrane, and on the inside by the wall and<br />

then the plasma membrane of the pathogen (Fig.<br />

5.21; Coffey & Wilson, 1983; Coffey & Gees, 1991).<br />

Haus<strong>to</strong>ria of Phy<strong>to</strong>phthora do not normally<br />

contain nuclei, although one may be situated<br />

near the branching point within the intercellular<br />

hypha (Fig. 5.21a).<br />

Asexual reproduction<br />

The sporangia of Phy<strong>to</strong>phthora spp. are usually<br />

pear-shaped or lemon-shaped (Fig. 5.22a) and<br />

arise on simple or branched sporangiophores<br />

which are more clearly differentiated than


104 STRAMINIPILA: OOMYCOTA<br />

Fig 5.20 Phy<strong>to</strong>phthorainfestans.<br />

(a) Sporangiophores penetrating a s<strong>to</strong>ma of a<br />

pota<strong>to</strong> leaf. (b) Zoospores and zoospore cysts, one<br />

formed inside a zoosporangium. (c) Intercellular<br />

mycelium from a pota<strong>to</strong> tuber showing the<br />

finger-like haus<strong>to</strong>ria penetrating the cell walls.<br />

Note the thickening of the cell walls around the<br />

haus<strong>to</strong>rium.<br />

those of Pythium. On the host plant, the sporangiophores<br />

may emerge through the s<strong>to</strong>mata,<br />

as in P. infestans (Fig. 5.20a). The first sporangium<br />

is terminal, but the hypha bearing it may push<br />

it <strong>to</strong> one side and form further sporangia by<br />

sympodial growth. Mature sporangia of most<br />

species have a terminal papilla which appears as<br />

a plug because it consists of material different<br />

from the sporangial wall (Coffey & Gees, 1991).<br />

In species of Phy<strong>to</strong>phthora infecting aerial<br />

plant organs, the sporangia are detached, possibly<br />

aided by hygroscopic twisting of the sporangiophore<br />

on drying, and are dispersed by wind<br />

before germinating. In aquatic or soil-borne<br />

forms, zoospore release commonly occurs<br />

whilst the sporangia are still attached; internal<br />

proliferation of attached sporangia may occur.<br />

Whether deciduous or not, sporangia may<br />

germinate either directly by means of a germ<br />

tube, or by releasing zoospores. The latter seems<br />

<strong>to</strong> be the original route because undifferentiated<br />

sporangia contain pre-formed flagella within<br />

their cy<strong>to</strong>plasm, and these are degraded under<br />

unfavourable conditions leading <strong>to</strong> direct germination<br />

(Hemmes, 1983; Erwin & Ribeiro, 1996).<br />

The mode of germination is dependent on<br />

environmental parameters. For example, in<br />

P. infestans, uninucleate zoospores are produced<br />

below 15°C whilst above 20°C multinucleate<br />

germ tubes arise. Further, with increasing age<br />

sporangia lose their capacity <strong>to</strong> produce zoospores<br />

and tend <strong>to</strong> germinate directly. In<br />

P. cac<strong>to</strong>rum, sporangia have been preserved for<br />

several months under moderately dry conditions.


PYTHIALES<br />

105<br />

Fig 5.21 TEM images of haus<strong>to</strong>ria of P. infestans. (a) Mature haus<strong>to</strong>rium within a leaf cell of pota<strong>to</strong>. (b) The basal region of a<br />

haus<strong>to</strong>rium.The haus<strong>to</strong>rium contains fungal vacuoles (FV) and mi<strong>to</strong>chondria (M) but no nuclei. However, a nucleus (NF) is located<br />

within the intercellular hypha close <strong>to</strong> the branch point.The plant <strong>to</strong>noplast (T), plant extrahaus<strong>to</strong>rial membrane (EM),<br />

extrahaus<strong>to</strong>rial matrix (EX) and fungal wall (FW) are visible.The seemingly empty space surrounding the haus<strong>to</strong>rium is the plant<br />

vacuole (V). Both images reprinted from Coffey and Wilson (1983) by copyright permission of the National Research Council of<br />

Canada.Original prints kindly provided by M.D.Coffey.<br />

When water becomes available again, such<br />

sporangia may germinate by the formation of<br />

a vegetative hypha, or a further sporangium.<br />

Thick-walled asexual spherical chlamydospores<br />

have also been described for many<br />

Phy<strong>to</strong>phthora spp. and can survive in soil for<br />

several years (Ribeiro, 1983; Erwin & Ribeiro,<br />

1996). The morphological differences between<br />

sporangia, chlamydospores and oospores are<br />

illustrated in Fig. 5.22.<br />

Once formed, mature sporangia may remain<br />

undifferentiated for several hours or even days,<br />

but zoospore differentiation can be induced by<br />

suspending mature sporangia in chilled water or<br />

soil extract. Detailed methods <strong>to</strong> trigger zoospore<br />

release have been established for many species<br />

(Erwin & Ribeiro, 1996). Once cold-shock has<br />

been received, differentiation can be completed<br />

in less than 60 min and probably involves<br />

cAMP-mediated signalling cascades (Yoshikawa<br />

& Masago, 1977). The processes of differentiation<br />

of sporangial pro<strong>to</strong>plasm in<strong>to</strong> zoospores differ<br />

in certain details between Phy<strong>to</strong>phthora and<br />

the Saprolegniales (see Hardham & Hyde, 1997).<br />

For instance, in Saprolegnia the central vacuole<br />

is prominent and its membrane as well as<br />

the plasma membrane contribute <strong>to</strong> the plasma<br />

membranes of the developing zoospores (p. 81).<br />

In contrast, in Phy<strong>to</strong>phthora the central vacuole<br />

disappears from the young sporangium before<br />

cleavage of the cy<strong>to</strong>plasm begins, and the<br />

plasma membrane remains intact even after<br />

zoospores have become fully differentiated. The<br />

zoospore plasma membranes therefore mostly<br />

originate from Golgi-derived cleavage cisternae<br />

(Hyde et al., 1991). Detailed cy<strong>to</strong>logical studies


106 STRAMINIPILA: OOMYCOTA<br />

Fig 5.22 Reproductive structures in Phy<strong>to</strong>phthora cac<strong>to</strong>rum. (a) Sporangia. (b) Chlamydospore. (c) Oospore showing the<br />

paragynous mode of fertilization. (d) Oospore with amphigynous fertilization. (b d) <strong>to</strong> same scale.<br />

have revealed an important role of microtubules<br />

in organizing the distribution of nuclei<br />

during zoospore formation (Hyde & Hardham,<br />

1992, 1993). Cleavage of the cy<strong>to</strong>plasm of a<br />

zoospore begins close <strong>to</strong> that end of the nucleus<br />

which subsequently points <strong>to</strong>wards the ventral<br />

groove. At this stage, three types of vesicle which<br />

become important during zoospore encystment<br />

also move in<strong>to</strong> their positions: large peripheral<br />

vesicles, dorsal vesicles, and small ventral vesicles.<br />

When the pre-formed flagella have been<br />

inserted, the zoospores acquire their mobility<br />

(Hardham, 1995). Zoospores are either discharged<br />

directly through the plug after this has dissolved,<br />

or they are transferred in<strong>to</strong> a very transient<br />

membranous vesicle which forms outside the<br />

opened plug upon discharge and bursts one or<br />

a few seconds later (Gisi, 1983). Since the plasma<br />

membrane of the sporangium has not become<br />

part of the zoospore membranes, the membranous<br />

vesicle is probably continuous with the<br />

plasma membrane.<br />

Encystment of zoospores<br />

Zoospores of Phy<strong>to</strong>phthora swim for several hours,<br />

travelling distances of a few centimetres in water<br />

or wet soil, although they can be spread much<br />

further by passive movement within water<br />

currents (Newhook et al., 1981). They are<br />

attracted chemotactically <strong>to</strong> plant roots by<br />

non-specific root exudates such as amino acids,<br />

host-specific substances, or the electrical field<br />

generated by plant roots (Carlile, 1983; Deacon<br />

& Donaldson, 1993; Tyler, 2002). No equivalent<br />

studies seem <strong>to</strong> have been carried out for zoospores<br />

of Phy<strong>to</strong>phthora infecting leaves. The<br />

process of zoospore encystment described below<br />

for Phy<strong>to</strong>phthora seems <strong>to</strong> apply also <strong>to</strong> Pythium<br />

(Hardham, 1995). It is an act of regulated secretion,<br />

i.e. the release of pre-formed contents by<br />

synchronous fusion of vesicles with the plasma<br />

membrane. Regulated secretion is common in<br />

animal cells, e.g. in epithelial or neuronal<br />

systems, but in fungi it is probably confined <strong>to</strong><br />

encysting zoospores.<br />

Zoospores of Phy<strong>to</strong>phthora are kidney-shaped;<br />

both flagella arise from the kine<strong>to</strong>some boss<br />

protruding from within the longitudinal groove<br />

at the ventral surface. The anterior end of the<br />

spore is indicated externally by the straminipilous<br />

flagellum and internally by the water<br />

expulsion vacuole; the nucleus is located in


PYTHIALES<br />

107<br />

Fig 5.23 Schematic drawings of a zoospore of Phy<strong>to</strong>phthora (not <strong>to</strong> scale). (a) Longitudinal section. (b) Transverse section of the<br />

anterior region showing several types of vesicle, namely the water-expulsion vacuole (WEV), fingerprint vacuole (FPV), large<br />

peripheral vesicles (LPV), small ventral vesicles (SVV), small dorsal vesicles (SDV) and peripheral cisternae (PC). Mi<strong>to</strong>chondria (Mit)<br />

with unusually lamellate cristae are also indicated. a modified from Dick (2001b); b based on the ultrastructural work of<br />

Hardham et al.(1991).<br />

the posterior half of the spore (Fig. 5.23a).<br />

The nucleus is associated with the microtubular<br />

roots of the flagella which force it in<strong>to</strong> a<br />

somewhat conical shape, the pointed end pointing<br />

<strong>to</strong>wards the kine<strong>to</strong>some boss. Zoospores<br />

contain several vesicular compartments. Their<br />

positions are drawn schematically in Fig. 5.23,<br />

and electron micrographs are provided in<br />

Fig. 5.24. Fingerprint vacuoles, equivalent <strong>to</strong> the<br />

dense-body vesicles of Saprolegnia and Achlya, are<br />

defined by the lamellate structure of their<br />

contents, presumably deposits of b-1,3-glucan<br />

(mycolaminarin) and phosphate. Fingerprint<br />

vacuoles are located mainly in the interior of<br />

the zoospore and play no part in the encystment<br />

process but are thought <strong>to</strong> provide carbon and<br />

energy reserves during subsequent germination<br />

of the cyst (Gubler & Hardham, 1990). In zoospores<br />

of Phy<strong>to</strong>phthora cinnamomi, there are<br />

several kinds of peripheral vesicle which have<br />

been distinguished morphologically (Fig. 5.23)<br />

and by labelling with specific antibodies. When<br />

zoospores approach a root, the groove of the<br />

ventral surface faces the root surface, initial<br />

contact presumably being made by the flagella.<br />

Attachment of the zoospore is achieved by means<br />

of a glue discharged by the synchronous fusion<br />

of the small ventral vesicles with the ventral<br />

plasma membrane (Hardham & Gubler, 1990).<br />

At the same time, the small dorsal vesicles<br />

also secrete their contents, leading <strong>to</strong> the<br />

deposition of the first cyst wall (Figs. 5.24c,d;<br />

Gubler & Hardham, 1988). The process of<br />

exocy<strong>to</strong>sis is complete within 2 min of receiving<br />

the encystment trigger. In contrast, the large<br />

peripheral vesicles do not fuse with the plasma<br />

membrane but withdraw <strong>to</strong> the centre of the<br />

cyst. Their contents are proteinaceous and<br />

probably serve as reserves for the germination<br />

process. Peripheral cisternae, ultrastructurally<br />

distinct from the ER, line the inside of the<br />

zoospore plasma membrane and disappear<br />

during encystment (Hardham et al., 1991;<br />

Hardham, 1995).


108 STRAMINIPILA: OOMYCOTA<br />

Fig 5.24 Ultrastructure of Phy<strong>to</strong>phthora cinnamomi zoospores as seen with theTEM. (a) Oblique section through a zoospore.<br />

Several kinds of vesicle are visible, as are mi<strong>to</strong>chondria, the water-expulsion vacuole (arrow) and the conical nucleus with its<br />

prominent nucleolus. (b) Fingerprint vacuoles. (c,d) Immunogold labelling of wall material located within dorsal vesicles before<br />

(c) and in the cyst wall1min after (d) encystment of the zoospore. (a,b) reproduced from Hardham and Hyde (1997), with permission<br />

from Elsevier; (c,d) previously unpublished work. All images kindly provided by F.Gubler and A.R. Hardham.<br />

Zoospore encystment can be triggered by<br />

several stimuli, e.g. contact with host cell surface<br />

polysaccharides, change in medium composition,<br />

or presence of root exudates. Commitment<br />

<strong>to</strong> encystment occurs within 20 30 s of receiving<br />

the stimulus (Paktitis et al., 1986). Complex<br />

signalling cascades involving Ca 2þ and phospholipase<br />

D are involved (Zhang et al., 1992), and<br />

commitment <strong>to</strong> several future developmental<br />

processes is made before the onset of encystment,<br />

including the point of germ tube emergence<br />

(Hardham & Gubler, 1990).<br />

Zoospore cysts germinate quite rapidly after<br />

their formation, usually by means of a germ tube<br />

which infects the plant roots directly. In the case<br />

of hard surfaces such as leaves, the germ tube<br />

may form an appressorium which mediates<br />

infection (see pp. 378 381).


PYTHIALES<br />

109<br />

Fig 5.25 Oogonial<br />

development in<br />

Phy<strong>to</strong>phthora.<br />

(a f) Stages of<br />

development in<br />

P. erythroseptica.<br />

(g i). Stages of<br />

development in<br />

P. cac<strong>to</strong>rum.<br />

Sexual reproduction<br />

Oospore formation is dependent on sterols and<br />

mating hormones (p. 95) and may be homo- or<br />

heterothallic. Phylogenetic studies have indicated<br />

that the former is ancestral, heterothallism<br />

having arisen repeatedly within the genus<br />

Phy<strong>to</strong>phthora (Kroon et al., 2004). Two distinct<br />

types of antheridial arrangement are found.<br />

In P. fragariae, P. megasperma and a number of<br />

other species, antheridia are attached laterally <strong>to</strong><br />

the oogonium and are described as paragynous<br />

meaning ‘beside the female’ (Figs. 5.22c, 5.25g i).<br />

In other Phy<strong>to</strong>phthora species such as P. infestans, P.<br />

cinnamomi and P. erythroseptica, the oogonium,<br />

during its development, penetrates and grows<br />

through the antheridium (Hemmes, 1983). The<br />

oogonial hypha emerges above the antheridium<br />

and inflates <strong>to</strong> form a spherical oogonium, with<br />

the antheridium persisting as a collar around its<br />

base (Figs. 5.25a f). This arrangement of the<br />

antheridium is termed amphigynous (‘around<br />

the female’). In some species (e.g. P. cac<strong>to</strong>rum,<br />

P. clandestina, P. medicaginis), both types of arrangement<br />

may be found (Figs. 5.22c,d); one or the<br />

other may predominate, depending on strain<br />

and culture conditions (Erwin & Ribeiro, 1996).<br />

Both the oogonia and antheridia contain<br />

several diploid nuclei, but as the oosphere<br />

matures only a single nucleus remains at the<br />

centre while the remaining nuclei are included<br />

in the periplasm, i.e. the space between the<br />

oosphere and the oogonial walls (see Fig. 5.2).<br />

Meiosis occurs in the antheridium and oogonium<br />

(Shaw, 1983). Fertilization tubes have been<br />

observed and a single haploid nucleus is introduced<br />

from the antheridium in<strong>to</strong> the oosphere<br />

(Fig. 5.26). Fusion between the oosphere nucleus<br />

and the antheridial nucleus is delayed. Even<br />

mature, dormant oospores may still be binucleate,<br />

karyogamy usually occurring after breakage<br />

of dormancy as a first step <strong>to</strong>wards germination<br />

(Jiang et al., 1989).<br />

Following fertilization, the physiology<br />

and ultrastructure of the oospore change <strong>to</strong>


110 STRAMINIPILA: OOMYCOTA<br />

Fig 5.26 Phy<strong>to</strong>phthora cac<strong>to</strong>rum.<br />

Development of oogonium, antheridium<br />

and oospore. (a) Initials of oogonium<br />

and antheridium. (b) Oogonium and<br />

antheridium grown <strong>to</strong> full size: the<br />

oogonium has about 24 nuclei and the<br />

antheridium about 9. (c) Development<br />

of a septum at the base of each, and<br />

degeneration of some nuclei in each until<br />

the oogonium has 8 or 9 nuclei and the<br />

antheridium 4 or 5. (d) A simultaneous<br />

division of the surviving nuclei in<br />

oogonium and antheridium.The<br />

pro<strong>to</strong>plast has large vacuoles.<br />

(e) Separation of oosphere from<br />

periplasm. Nuclei divide in the periplasm<br />

prior <strong>to</strong> degeneration.The oogonium<br />

presses in<strong>to</strong> the antheridium. (f) Entry of<br />

one antheridial nucleus by a fertilization<br />

tube.The pro<strong>to</strong>plasm and remaining<br />

nuclei of the antheridium degenerate.<br />

(g) Development of oospore wall.<br />

(h) The oospore enters its dormant<br />

period with exospore formed from dead<br />

periplasm, endospore deposited inside<br />

it, and paired nuclei in association but<br />

not yet fused. (a h) are composite<br />

drawings of eight stages in sequence<br />

(after Blackwell,1943).<br />

a resting state. Oospore differentiation proceeds<br />

from the outside inwards (centripetal development).<br />

The oospore has a thin outer wall<br />

(epispore) which is derived from the periplasm<br />

and appears <strong>to</strong> consist of pectic substances. The<br />

inner oospore wall (endospore) is rich in b-1,3-<br />

glucans which form a major s<strong>to</strong>rage reserve and<br />

are mobilized by glucanases just prior <strong>to</strong> germination<br />

(Erwin & Ribeiro, 1996). Within the<br />

developing oospore, the numerous small lipid<br />

droplets coalesce in<strong>to</strong> a few large ones. Lipids are<br />

undoubtedly the major endogenous s<strong>to</strong>rage<br />

reserve in the spores of Oomycota (Dick, 1995)<br />

and many other fungi. Later, the dense body<br />

vesicles which are rich in mycolaminarin and<br />

phosphate fuse <strong>to</strong>gether, giving one large structure,<br />

the ooplast. Like the endospore, the ooplast<br />

is consumed during germination whereas some<br />

lipid droplets are saved and are translocated in<strong>to</strong><br />

the germ tube (Hemmes, 1983). Considering their<br />

thick walls and abundant s<strong>to</strong>rage reserves, it is<br />

not surprising that oospores are the longest-lived


PYTHIALES<br />

111<br />

propagule of Phy<strong>to</strong>phthora, being capable of<br />

surviving in soil for many years.<br />

5.3.4 Phy<strong>to</strong>phthora infestans,causeof<br />

pota<strong>to</strong> late blight<br />

Late blight of pota<strong>to</strong> caused by P. infestans is a<br />

no<strong>to</strong>rious disease. In the period between 1845<br />

and 1848 it resulted in famine across much of<br />

Europe, and especially in Ireland where most<br />

people had come <strong>to</strong> depend on the pota<strong>to</strong> as<br />

their major source of food. In Ireland alone, the<br />

population size dropped from over 8 million in<br />

1841 <strong>to</strong> 6.5 million in 1851 (Salaman, 1949). The<br />

his<strong>to</strong>ry of the Great Famine has been ably<br />

documented by Large (1940), Woodham-Smith<br />

(1962) and Schumann (1991). The social and<br />

political repercussions of this tragedy have<br />

been immense and still reverberate <strong>to</strong>day.<br />

An enormous amount of literature about<br />

P. infestans has been published over the past<br />

150 years, including several books (Ingram &<br />

Williams, 1991; Lucas et al., 1991; Dowley et al.,<br />

1995). It has been estimated that about 10% of<br />

the entire phy<strong>to</strong>pathological literature is concerned<br />

just with this one species. None the less,<br />

there are many uncomfortable gaps in our<br />

knowledge, and the fungus continues <strong>to</strong> provide<br />

unpleasant surprises <strong>to</strong> this day.<br />

Origin and spread<br />

The probable centre of evolution of most Solanum<br />

spp. and hence also their pathogens, notably<br />

P. infestans, lies in Mexico (Niederhauser, 1991),<br />

although the pota<strong>to</strong> (S. tuberosum) was first<br />

cultivated in South America. There are several<br />

theories accounting for the spread of P. infestans<br />

round the world (Ristaino, 2002). In the early<br />

1840s P. infestans rapidly spread <strong>to</strong> North America,<br />

and it is generally assumed that it was introduced<br />

<strong>to</strong> Europe (Belgium) in June 1845 with a<br />

shipment of contaminated pota<strong>to</strong>es (Bourke,<br />

1991). Phy<strong>to</strong>phthora infestans is heterothallic, and<br />

there is good evidence that in the first wave<br />

of migration in 1845 only the A1 mating type<br />

reached Europe (Goodwin et al., 1994a). Over the<br />

next century or more, the fungus probably<br />

survived entirely on an asexual life cycle, overwintering<br />

in tubers infected during the previous<br />

season and discarded <strong>to</strong>gether with shoots and<br />

other debris in the field. Despite the absence of<br />

sexual reproduction, P. infestans showed a considerable<br />

genetic adaptability, as documented by<br />

its ability <strong>to</strong> break the resistance bred in<strong>to</strong> new<br />

pota<strong>to</strong> cultivars (p. 114), and also the rapid<br />

emergence of strains resistant against newly<br />

introduced fungicides (p. 112).<br />

A second wave of P. infestans migration<br />

brought the A2 mating type from central<br />

Mexico <strong>to</strong> North America and Europe where it<br />

was first isolated in 1981 (Hohl & Iselin, 1984).<br />

It is now established worldwide (Spielman et al.,<br />

1991; Fry et al., 1993; Gillis, 1993; Goodwin et al.,<br />

1994b). The enhanced genetic recombination<br />

brought about by sexual reproduction is catalysing<br />

a change in the genetic make up of<br />

P. infestans, which may be leading <strong>to</strong> an explosive<br />

evolution of new P. infestans strains (Fry et al.,<br />

1993; Goodwin et al., 1995). This situation is seen<br />

as the biggest threat posed by P. infestans since<br />

the 1840s (Fry & Goodwin, 1997).<br />

Epidemiology<br />

There is clear genetic evidence of sexual reproduction<br />

taking place in the field, and it is also<br />

possible that oospores contribute <strong>to</strong> the survival<br />

of P. infestans in soil during the winter (Andrivon,<br />

1995). Additionally, the fungus has a good<br />

capacity <strong>to</strong> survive the winter without oospores.<br />

A very low proportion of infected tubers left on<br />

the field gives rise <strong>to</strong> infected ‘volunteer’ plants<br />

in the following spring. In experimental plots,<br />

the proportion of infected plants developing<br />

from naturally or artificially infected tubers<br />

was found <strong>to</strong> be less than 1% (Hirst & Stedman,<br />

1960). Nevertheless, such infected shoots form<br />

foci within the crop from which the disease<br />

spreads. The sporangia of P. infestans are deciduous,<br />

and they are blown from diseased shoots<br />

<strong>to</strong> healthy leaves where they germinate either<br />

by the formation of germ tubes or zoospores.<br />

Zoospore production is favoured by lower<br />

temperatures (9 15°C). After swimming for a<br />

time, the zoospores encyst and then form germ<br />

tubes which usually penetrate the epidermal<br />

walls of the pota<strong>to</strong> leaf, or occasionally enter the<br />

s<strong>to</strong>mata. An appressorium is formed at the tip<br />

of the germ tube, attaching the zoospore cyst


112 STRAMINIPILA: OOMYCOTA<br />

firmly <strong>to</strong> the leaf. Penetration of the cell wall is<br />

probably achieved by a combination of mechanical<br />

and enzymatic action and can occur within<br />

2 h. Within the leaf tissue, an intercellular<br />

mycelium develops and haus<strong>to</strong>ria are formed<br />

where hyphae contact host cell walls (Fig. 5.21).<br />

The resulting lesion acquires a dark green watersoaked<br />

appearance associated with tissue disintegration<br />

(Plate 2e). Such lesions are visible<br />

within 3 5 days of infection under suitable conditions<br />

of temperature and humidity. Around<br />

the margin of the advancing lesion on the lower<br />

surface of the leaf, a zone of sporulation is found<br />

in which sporangiophores emerge through the<br />

s<strong>to</strong>mata (Fig. 5.20a). Sporulation is most prolific<br />

during periods of high humidity and commonly<br />

occurs at night following the deposition of dew.<br />

In pota<strong>to</strong> crops, as the leaf canopy closes over<br />

between the rows <strong>to</strong> cover the soil, a humid<br />

microclimate is established which may result<br />

in extensive sporulation. As the foliage dries<br />

during the morning, the sporangiophore undergoes<br />

hygroscopic twisting which results in the<br />

flicking-off of sporangia. Thus the concentration<br />

of sporangia in the air usually shows a<br />

characteristic diurnal fluctuation, with a peak<br />

around 10 a.m. Although sporangia can survive<br />

drying if they are rehydrated slowly (Minogue &<br />

Fry, 1981), in practice the long-range spread of<br />

inoculum is probably by sporangia in contact<br />

with water drops (Warren & Colhoun, 1975).<br />

The destructive action of P. infestans is directly<br />

associated with the killing of pho<strong>to</strong>synthetically<br />

active foliage. When about 75% of the leaf tissue<br />

has been destroyed, further increase in the<br />

weight of the crop ceases (Cox & Large, 1960).<br />

Thus, the earlier the onset of the epidemic, the<br />

more serious the consequences. To a certain<br />

extent, the crop reduction may be offset by the<br />

fact that epidemics are more common in rainy<br />

cool seasons which are conducive <strong>to</strong> higher crop<br />

yields.<br />

Phy<strong>to</strong>phthora infestans can also cause severe<br />

post-harvest crop losses because tubers can be<br />

infected by sporangia falling on<strong>to</strong> them, either<br />

during growth or lifting. Such infected tubers<br />

may rot in s<strong>to</strong>rage, and the diseased tissue is<br />

susceptible <strong>to</strong> secondary bacterial and fungal<br />

infections.<br />

Chemical control<br />

By spraying with suitable fungicides, epidemic<br />

spread of the disease can be delayed. This results<br />

in a prolongation of pho<strong>to</strong>synthetic activity of<br />

the pota<strong>to</strong> foliage and hence an increase in yield.<br />

<strong>Fungi</strong>cides developed against the Eumycota are<br />

often ineffective against Oomycota such as<br />

Phy<strong>to</strong>phthora because the latter differ in fundamental<br />

biochemical principles, including many<br />

of the molecular targets of fungicides active<br />

against Eumycota (Bruin & Edging<strong>to</strong>n, 1983;<br />

Griffith et al., 1992). In 1991, about 20% of the<br />

<strong>to</strong>tal amount of money spent on chemicals for<br />

controlling plant diseases worldwide was used<br />

for the control of Oomycota (Schwinn & Staub,<br />

1995).<br />

The first of all fungicides was Bordeaux<br />

mixture, an inorganic formulation containing<br />

copper sulphate and calcium oxide which was<br />

found <strong>to</strong> be effective against downy mildew<br />

of vines caused by Plasmopara viticola, another<br />

member of the Oomycota (see p. 119; Large, 1940;<br />

Erwin & Ribeiro, 1996). Oomycota in general are<br />

extremely sensitive <strong>to</strong> copper ions, and Bordeaux<br />

mixture is still widely used (Agrios, 2005).<br />

The dithiocarbamates such as zineb or<br />

maneb (Fig. 5.27a) were among the first organic<br />

fungicides <strong>to</strong> be developed. They act against<br />

a wide range of fungi, including Oomycota,<br />

because of their non-selective mode of action.<br />

The molecule is sufficiently apolar <strong>to</strong> diffuse<br />

across the fungal plasma membrane; once inside,<br />

it is metabolized, and the released isothiocyanate<br />

radical (Fig. 5.27b) reacts with the sulphydryl<br />

groups of amino acids (Agrios, 2005).<br />

The most important agrochemicals against<br />

Oomycota are the phenylamides such as metalaxyl<br />

(Fig. 5.27c) which are systemic fungicides,<br />

i.e. they can enter the plant and are translocated<br />

throughout it. Metalaxyl appears <strong>to</strong> inhibit the<br />

transcription of ribosomal RNA in Oomycota<br />

but not Eumycota (Davidse et al., 1983). This is<br />

an inhibition of a specific biochemical target,<br />

and the immense genetic variability of P. infestans<br />

enabled it <strong>to</strong> develop resistance against metalaxyl<br />

in the early 1980s shortly after this was<br />

released for agricultural use (Davidse et al.,<br />

1991). Resistance is now widespread and has<br />

serious implications for future control of


PYTHIALES<br />

113<br />

Fig 5.27 <strong>Fungi</strong>cides against P. infestans. (a) The dithiocarbamate maneb which is active against Oomycota and Eumycota. (b) The<br />

isothiocyanate radical released by metabolism of dithiocarbamates by fungal hyphae. (c) The phenylamide metalaxyl which is active<br />

only against Oomycota. (d) Aluminium ethyl phosphonate (fosetyl-Al). (e) Cyazofamid, a new fungicide specific against Oomycota.<br />

(f) Famoxadone, a new fungicide active against Oomycota and Eumycota.<br />

Phy<strong>to</strong>phthora spp. (Erwin & Ribeiro, 1996).<br />

Phenylamides are now protected by being used<br />

in a cocktail, e.g. with the less-specific dithiocarbamates,<br />

and tailor-made application regimes<br />

are recommended for each year and each region<br />

(Staub, 1991).<br />

The phosphonates are a different type of<br />

fungicide against Phy<strong>to</strong>phthora spp. Fosetyl Al<br />

(aluminium ethyl phosphonate; Fig. 5.27d) is<br />

readily taken up by plants in which it is broken<br />

down <strong>to</strong> release phosphorous acid (¼ phosphonate),<br />

which seems <strong>to</strong> be the active principle<br />

(Griffith et al., 1992). Fosetyl Al as well as phosphorous<br />

acid can move downwards through the<br />

phloem and upwards in the xylem, showing<br />

similar transport characteristics as sucrose<br />

(Ouimette & Coffey, 1990; Erwin & Ribeiro,<br />

1996). The mode of action of phosphonates is<br />

not known but is likely <strong>to</strong> be complex, with a<br />

stimula<strong>to</strong>ry effect also on the host plant immune<br />

system (Molina et al., 1998). Although active<br />

only against pota<strong>to</strong> tuber blight but not foliar<br />

blight caused by P. infestans (L. R. Cooke & Little,<br />

2002), phosphonates are effective against a wide<br />

range of root-infecting Phy<strong>to</strong>phthora spp. and<br />

even show good curative properties (Erwin &<br />

Ribeiro, 1996).<br />

A useful introduction <strong>to</strong> current fungicides<br />

and their modes of action has been provided<br />

by Uesugi (1998). Because of the enormous<br />

economic significance of P. infestans and other<br />

Oomycota, new fungicide candidates are continually<br />

being developed and introduced in<strong>to</strong> the<br />

market. Two recent examples are cyazofamid<br />

(Fig. 5.27e) and famoxadone (Fig. 5.27f). Both<br />

inhibit mi<strong>to</strong>chondrial respiration. However,<br />

whilst the former is specific against Oomycota<br />

(Sternberg et al., 2001), famoxadone inhibits<br />

both Oomycota and Eumycota (Mitani et al.,<br />

2002). Its molecular target is different from<br />

that of cyazofamid but probably the same as<br />

that of the strobilurins (see Figs. 13.15e,f),<br />

as indicated by the development of crossresistance<br />

in fungal pathogens against famoxadone<br />

and strobilurins.<br />

Disease forecast<br />

To avoid unnecessary spraying and <strong>to</strong> ensure<br />

that timely spray applications are made, it has<br />

proven possible <strong>to</strong> provide forecasts of the<br />

incidence of pota<strong>to</strong> blight epidemics for certain<br />

countries. Beaumont (1947) analysed the incidence<br />

of blight epidemics in south Devon<br />

(England) and established that a ‘temperature<br />

humidity rule’ controls the relationship between<br />

blight epidemics and weather. After a certain<br />

date (which varies with the locality) and assuming<br />

that inoculum on volunteer plants is always


114 STRAMINIPILA: OOMYCOTA<br />

present, Beaumont (1947) predicted that blight<br />

would follow within 15 22 days of a period<br />

of at least 48 h during which the minimum temperature<br />

was not less than 10°C and the relative<br />

humidity was over 75%. The warm humid<br />

weather during this Beaumont period provides<br />

conditions suitable for sporulation and the<br />

initiation of new infections. Modified in the<br />

light of experience and adapted <strong>to</strong> regional<br />

climates, computerized forecasting systems are<br />

now used worldwide, limiting fungicide applications<br />

<strong>to</strong> situations in which they are necessary<br />

(Doster & Fry, 1991; Erwin & Ribeiro, 1996). After<br />

receipt of a blight warning, fungicide sprays are<br />

applied prophylactically by the farmer, irrespective<br />

of whether P. infestans is actually present in<br />

his field or not.<br />

Haulm destruction<br />

The danger of infection of tubers by sporangia<br />

falling on<strong>to</strong> them from foliage at lifting time can<br />

be minimized by ensuring that all the foliage is<br />

destroyed before lifting. This is achieved by<br />

spraying the foliage with herbicides 2 3 weeks<br />

before harvest time. The ridging of pota<strong>to</strong> tubers<br />

also helps <strong>to</strong> protect the tubers from infection.<br />

Although sporangia may survive in the soil<br />

for several weeks, they do not penetrate deeply<br />

in<strong>to</strong> it.<br />

Crop sanitation<br />

In principle, one infected volunteer plant per<br />

hectare is sufficent <strong>to</strong> initiate an epidemic.<br />

This is because late blight is a typical multicyclic<br />

disease, with numerous cycles of reproduction<br />

occurring in a single growing season under<br />

favourable conditions, leading <strong>to</strong> the rapid<br />

build up of inoculum. Crop sanitation, which is<br />

effective against single-cycle diseases, therefore<br />

has only limited value in the control of<br />

P. infestans (van der Plank, 1963).<br />

Breeding for major gene resistance<br />

A worldwide screening of Solanum spp. showed<br />

that a number of them have natural resistance <strong>to</strong><br />

P. infestans. One species which has proven <strong>to</strong> be an<br />

important source of resistance is S. demissum<br />

which grows in Mexico, the presumed centre<br />

of origin of P. infestans. Although this species is<br />

valueless in itself for commercial cultivation,<br />

it is possible <strong>to</strong> cross it with S. tuberosum, and<br />

some of the progeny are resistant <strong>to</strong> the disease.<br />

Solanum demissum contains at least four major<br />

genes for resistance (R 1 , R 2 , R 3 and R 4 ), <strong>to</strong>gether<br />

with a number of minor genes which determine<br />

the degree of susceptibility in susceptible<br />

varieties (Black, 1952). The four genes may be<br />

absent from a particular host strain, or they may<br />

be present singly (e.g. R 1 ), in pairs, in threes,<br />

or all <strong>to</strong>gether, so that 16 host genotypes are<br />

possible representing different combinations of<br />

R genes. The identification of the R gene complex<br />

was dependent on the discovery that the fungus<br />

itself exists in a number of strains or physiological<br />

races. For each host R gene, the pathogen<br />

was assumed <strong>to</strong> carry a gene which enables it<br />

<strong>to</strong> overcome the effect of the R gene. This is<br />

the basis of the gene-for-gene hypothesis, and<br />

gene-for-gene interactions are common in<br />

many host pathogen interactions (Flor, 1971).<br />

Assuming a gene-for-gene situation for the<br />

interaction of P. infestans with S. tuberosum, 16<br />

races of P. infestans should theoretically be<br />

demonstrable. If the corresponding genes of<br />

the fungus are termed 1, 2, 3 and 4, then the<br />

different races can be labelled (0), (1), (2), etc.,<br />

(1.1), (1.2), etc., (1.2.3), (1.2.4), etc., and (1.2.3.4).<br />

By 1953, 13 of the 16 races had been identified,<br />

the prevalent race being Race 4. By 1969, 11 R<br />

genes had been recognized in Britain<br />

(Malcolmson, 1969). Resistance based on a<br />

small number of defined genes of major effect<br />

has been termed major gene resistance or<br />

race-specific resistance. Because of the uncanny<br />

ability of P. infestans <strong>to</strong> break major gene resistance<br />

even before the arrival of the A2 mating<br />

type in Europe and North America, attempts at<br />

breeding fully resistant pota<strong>to</strong> cultivars have<br />

now been abandoned (Wastie, 1991).<br />

The origin of physiological races is difficult <strong>to</strong><br />

determine. The occurrence and spread of resistance<br />

genes before the arrival of the A2 mating<br />

type may have been due <strong>to</strong> mutation followed<br />

by selection imposed by the monoculture of<br />

a resistant host. Another possibility is that<br />

the mycelium of P. infestans is heterokaryotic,<br />

carrying nuclei of more than one race. Yet<br />

another scenario is vegetative hybridization


PERONOSPORALES<br />

115<br />

followed by parasexual recombination (see<br />

p. 230); by mixing sporangia of two different<br />

races, new races with a different pattern of<br />

virulence <strong>to</strong>wards pota<strong>to</strong> varieties have been<br />

obtained after several cycles of inoculation<br />

(Malcolmson, 1970). The parasexual cycle has<br />

been experimentally demonstrated for P. parasitica<br />

using fungicide resistance as a genetic<br />

marker (Gu & Ko, 1998).<br />

Within 1 2 days of infection, tissues of resistant<br />

hosts undergo necrosis so rapidly that<br />

sporulation and further growth of the fungus<br />

cannot occur. Such a reaction is sometimes<br />

termed hypersensitivity, and the function of<br />

the R genes is <strong>to</strong> accelerate this host reaction.<br />

When pota<strong>to</strong> tubers are inoculated with an<br />

avirulent race of P. infestans, they respond by<br />

secreting antifungal substances called phy<strong>to</strong>alexins.<br />

Two of the phy<strong>to</strong>alexins formed by<br />

resistant tubers are rishitin and phytuberin.<br />

Rishitin, originally isolated from the pota<strong>to</strong><br />

variety Rishiri, is a bicyclic sesquiterpene.<br />

Tomiyama et al. (1968) showed that R 1 tuber<br />

tissue inoculated with an avirulent race of<br />

P. infestans produced over 270 times the amount<br />

of rishitin than when inoculated with a virulent<br />

race. The R genes of the pota<strong>to</strong> probably determine<br />

the ability of host tissue <strong>to</strong> recognize<br />

and respond <strong>to</strong> avirulent races of P. infestans<br />

(Day, 1974). The detailed molecular interactions<br />

which determine race specificity are, however,<br />

complex and still only incompletely unders<strong>to</strong>od<br />

at present (Friend, 1991).<br />

Breeding for field resistance<br />

In addition <strong>to</strong> the major genes for resistance<br />

in pota<strong>to</strong>, numerous other genes also exist<br />

which, although individually of small effect,<br />

may contribute <strong>to</strong> resistance if present <strong>to</strong>gether.<br />

Resistance of this kind is known as general<br />

resistance or field resistance, and some pota<strong>to</strong><br />

breeding programmes aim at producing varieties<br />

possessing it (Niederhauser, 1991). This is<br />

preferable <strong>to</strong> single-gene resistance because<br />

P. infestans is less likely <strong>to</strong> overcome the combined<br />

resistance of numerous minor genes<br />

simultaneously. Field resistance retards the<br />

infection process, e.g. by production of a particularly<br />

thick cuticle or by a leaf architecture<br />

unfavourable <strong>to</strong> infection, lowers the number of<br />

sporangia produced, and extends the time<br />

needed by the pathogen <strong>to</strong> initiate new infections<br />

(Wastie, 1991). Field resistance is equally<br />

effective against all physiological races of<br />

P. infestans, and it reduces the severity of an<br />

epidemic and consequently the need <strong>to</strong> apply<br />

fungicides (Erwin & Ribeiro, 1996).<br />

Toma<strong>to</strong> late blight<br />

P. infestans also causes significant worldwide crop<br />

losses of <strong>to</strong>ma<strong>to</strong> (Lycopersicon esculentum) which,<br />

like pota<strong>to</strong>, belongs <strong>to</strong> the Solanaceae. The<br />

general principles of control of <strong>to</strong>ma<strong>to</strong> late<br />

blight are similar <strong>to</strong> those described above for<br />

pota<strong>to</strong>, including fungicides used and blight<br />

forecasting (Erwin & Ribeiro, 1996). Many strains<br />

of P. infestans are capable of infecting both<br />

<strong>to</strong>ma<strong>to</strong> and pota<strong>to</strong>. However, since the resistance<br />

gene systems are different in these two hosts,<br />

correlations between virulence of a given strain<br />

on pota<strong>to</strong> and <strong>to</strong>ma<strong>to</strong> cannot be drawn (Legard<br />

et al., 1995).<br />

5.4 Peronosporales<br />

The Peronosporales are obligately biotrophic<br />

pathogens of a few groups of higher plants<br />

and are responsible for diseases mainly of<br />

aerial plant organs known collectively as<br />

downy mildews. The order currently comprises<br />

two families, the Peronosporaceae (Peronospora,<br />

Plasmopara, Bremia) and Albuginaceae (Albugo).<br />

There are about 250 species (Kirk et al., 2001).<br />

DNA sequencing data (Cooke et al., 2000;<br />

Riethmüller et al., 2002) are confusing at present<br />

because species of Phy<strong>to</strong>phthora (Pythiales) and<br />

Peronospora (Peronosporales) seem <strong>to</strong> intergrade<br />

in phylogenetic analyses. Peronospora seems<br />

more closely related <strong>to</strong> Phy<strong>to</strong>phthora than <strong>to</strong><br />

other members of the Peronosporales such as<br />

Albugo, which in turn may have affinity with<br />

Pythium. Considerable rearrangements between<br />

the Peronosporales and Pythiales will therefore<br />

have <strong>to</strong> be carried out at some point in the<br />

future. However, we prefer <strong>to</strong> retain the conventional<br />

system for the time being because the<br />

downy mildews (Peronosporales) represent a


116 STRAMINIPILA: OOMYCOTA<br />

convincing biological entity (Dick, 2001a). The<br />

key features distinguishing them from the<br />

Pythiales are as follows.<br />

First, they are obligate biotrophs and cannot<br />

be grown apart from their living host. The<br />

mycelium in the host tissues is coenocytic and<br />

intercellular, with haus<strong>to</strong>ria of various types<br />

penetrating the cell walls. No member of the<br />

Peronosporales has as yet been grown in axenic<br />

culture, although some can be propagated in<br />

dual culture with callus tissues of their plant<br />

hosts. None the less, some species (e.g. Plasmopara<br />

viticola) can cause cell damage <strong>to</strong> their hosts<br />

which leads <strong>to</strong> the leakage of cy<strong>to</strong>plasm (Lafon &<br />

Bulit, 1981). This is similar <strong>to</strong> the rots caused, for<br />

example, by Phy<strong>to</strong>phthora erythroseptica (Plate 2f)<br />

and suggests an incomplete adaptation <strong>to</strong> the<br />

biotrophic habit, tying in with the likely origin<br />

of Peronosporales from within the Pythiales<br />

(Dick, 2001a).<br />

Second, whereas Pythium and Phy<strong>to</strong>phthora spp.<br />

are typically able <strong>to</strong> attack a very wide range of<br />

host plants, Dick (2001a) has pointed out that<br />

Peronosporales parasitize a narrow range of<br />

angiosperm families, usually dicotyledons, and<br />

especially herbaceous plants which are either<br />

highly evolved or accumulate large amounts of<br />

secondary metabolites such as essential oils or<br />

alkaloids. Any one species of downy mildew is<br />

specific <strong>to</strong> only one or a few related host genera.<br />

Dick (2001a, 2002) has speculated that a coevolution<br />

of the downy mildews with herbaceous<br />

angiosperms occurred mainly in the Tertiary<br />

period, and as several independent events,<br />

whereby Phy<strong>to</strong>phthora and downy mildews share<br />

common ances<strong>to</strong>rs. The Peronosporaceae are<br />

relatively recent; Peronospora, along with its<br />

host plants, may have arisen in the mid <strong>to</strong> late<br />

Tertiary in the vicinity of Armenia and Iran.<br />

Plasmopara is probably of South American origin<br />

and dates back <strong>to</strong> the early Tertiary, whereas<br />

Bremia lactucae is a central European species.<br />

In contrast, the Albuginaceae (Albugo) are more<br />

ancient, with a late Cretaceous origin possibly in<br />

South America (Dick, 2002).<br />

A third major feature of the Peronosporales<br />

is the tendency of their sporangia <strong>to</strong> germinate<br />

directly, rather than by releasing zoospores.<br />

Many species have lost the ability <strong>to</strong> produce<br />

zoospores al<strong>to</strong>gether, their sporangia being functional<br />

‘conidia’ which are disseminated by wind.<br />

The sporangiophores are well-differentiated,<br />

showing determinate growth and branching<br />

patterns which provide characteristic features<br />

for identification. The production of directly<br />

germinating sporangia on well-defined sporangiophores<br />

represents an adaptation <strong>to</strong> the terrestrial<br />

lifestyle and supports the postulated origin<br />

of the Peronosporales in the drier Tertiary period<br />

(Dick, 2002). The life cycle of Peronosporales is<br />

similar <strong>to</strong> that of Phy<strong>to</strong>phthora (see Fig. 5.19).<br />

Sporangia infect directly or produce infective<br />

zoospores, leading <strong>to</strong> a new crop of sporangiophores<br />

and sporangia, and this asexual cycle<br />

spreads the disease during the vegetation period.<br />

Sexual reproduction is by means of oospores<br />

which are formed within the host tissue and<br />

survive adverse conditions after host death.<br />

Peronosporales cause economically significant<br />

diseases, and one of them Plasmopara<br />

viticola has had a major impact on agriculture<br />

and plant pathology because it led <strong>to</strong> the discovery<br />

of Bordeaux mixture (see p. 119). Overviews<br />

of the Peronosporales have been given by Spencer<br />

(1981), Smith et al. (1988) and Dick (2002).<br />

5.4.1 Peronospora (Peronosporaceae)<br />

Peronospora destruc<strong>to</strong>r causes a serious disease of<br />

onions and shallots whilst P. farinosa causes<br />

downy mildew of sugar beet, beetroot and<br />

spinach, but can also be found on weeds such<br />

as Atriplex and Chenopodium. Peronospora tabacina<br />

causes blue mould of <strong>to</strong>bacco. This name refers<br />

<strong>to</strong> the bluish purple colour of the sporangia,<br />

which is actually a feature of many species<br />

of Peronospora. Crop losses associated with<br />

P. tabacina can be up <strong>to</strong> 95%. This species was<br />

introduced in<strong>to</strong> Europe in 1958 and has spread<br />

rapidly since (Smith et al., 1988).<br />

Peronospora parasitica attacks members of the<br />

Brassicaceae. Although many specific names<br />

have been applied <strong>to</strong> forms of this fungus on<br />

different host genera, it is now cus<strong>to</strong>mary <strong>to</strong><br />

regard them all as belonging <strong>to</strong> a single species<br />

(Dickinson & Greenhalgh, 1977; Kluczewski &<br />

Lucas, 1983). Turnips, swede, cauliflower,<br />

Brussels sprouts and wallflowers (Cheiranthus)


PERONOSPORALES<br />

117<br />

Fig 5.28 Peronospora parasitica on Capsella bursa-pas<strong>to</strong>ris.<br />

(a) Sporangiophore. (b) Sporangium germinating by means<br />

of a germ tube. (c) L.S. of host stem showing intercellular<br />

mycelium and coarse lobed haus<strong>to</strong>ria.<br />

are commonly attacked, and the fungus is found<br />

particularly frequently on shepherd’s purse<br />

(Capsella bursa-pas<strong>to</strong>ris). Diseased plants stand<br />

out by their swollen and dis<strong>to</strong>rted stems bearing<br />

a white ‘fur’ of sporangiophores (Plate 2g). On<br />

leaves the fungus is associated with yellowish<br />

patches on the upper surface and the formation<br />

of white sporangiophores beneath. Sections of<br />

diseased tissue show a coenocytic intercellular<br />

mycelium and branched lobed haus<strong>to</strong>ria in<br />

certain host cells (Fig. 5.28c; Fraymouth, 1956).<br />

Following penetration of the host cell by<br />

P. parasitica, reactions are set up between the<br />

host pro<strong>to</strong>plasm and the invading fungus. The<br />

haus<strong>to</strong>rium becomes ensheathed by a layer of<br />

callose which is visible as a thickened collar<br />

around the haus<strong>to</strong>rial base in susceptible host<br />

plants, whereas the entire haus<strong>to</strong>rium may be<br />

coated by thick callose deposits in interactions<br />

showing a resistance response (Donofrio &<br />

Delaney, 2001). The general appearance of haus<strong>to</strong>ria<br />

of Peronospora is very similar <strong>to</strong> that of<br />

Phy<strong>to</strong>phthora shown in Fig. 5.21; the main body of<br />

the haus<strong>to</strong>rium is surrounded by host cy<strong>to</strong>plasm,<br />

the host plasma membrane, an extrahaus<strong>to</strong>rial<br />

matrix, the fungus cell wall, and the fungal<br />

plasma membrane (Fig. 5.29). Although the<br />

haus<strong>to</strong>ria undoubtedly play a major role in the<br />

nutrient uptake of the fungus from the host<br />

plant, it should be noted that intercellular<br />

hyphae are also capable of assimilating nutrients<br />

in planta (Clark & Spencer-Phillips, 1993; Spencer-<br />

Phillips, 1997).<br />

The sporangiophores emerge singly or in<br />

groups from s<strong>to</strong>mata. There is a s<strong>to</strong>ut main axis<br />

which branches dicho<strong>to</strong>mously <strong>to</strong> bear eggshaped<br />

sporangia at the tips of incurved<br />

branches (Fig. 5.28a). Detachment of sporangia<br />

is possibly caused by hygroscopic twisting of the<br />

sporangiophores related <strong>to</strong> changes in humidity.


118 STRAMINIPILA: OOMYCOTA<br />

Fig 5.29 Peronospora manshurica. Diagram of host pathogen interface in the haus<strong>to</strong>rial region.Fungal cy<strong>to</strong>plasm (FC) is bounded<br />

by the fungal plasma membrane (FP), lomasomes (LO) and the fungal cell wall (FW) in both the intercellular hyphae (right) and the<br />

haus<strong>to</strong>rium (centre).The relative positions of the host cell vacuole (V), host cy<strong>to</strong>plasm (HC) and host plasmalemma (HP) are<br />

indicated.The host cell wall (HW) terminates in a sheath (S).The zone of apposition (Z) separates the haus<strong>to</strong>rium from the host<br />

plasmalemma. Invaginations of the host plasmalemma and vesicular host cy<strong>to</strong>plasm are considered evidence for host secre<strong>to</strong>ry<br />

activity (sec). After Pey<strong>to</strong>n and Bowen (1963).<br />

In P. tabacina, however, it has been suggested that<br />

changes in turgor pressure of the sporangiophores<br />

occur which parallel changes in the water<br />

content of the <strong>to</strong>bacco leaf. Sporangia may be<br />

discharged actively by application of energy at<br />

their point of attachment <strong>to</strong> the sporangiophore.<br />

In the Sclerosporaceae (see Section 5.5), violent<br />

sporangial discharge also occurs. Upon alighting<br />

on a suitable host, sporangia of P. parasitica<br />

germinate by the formation of a germ tube<br />

rather than zoospores. The germ tube penetrates<br />

the wall of the epidermis by means of an<br />

appressorium (Fig. 5.28b).<br />

Oospores of P. parasitica, like those of most<br />

other Peronosporales, are embedded in senescent<br />

leaf tissues and are found throughout the season.<br />

There is evidence that some strains of the fungus<br />

are heterothallic whilst others are homothallic<br />

(McMeekin, 1960). Both the antheridium and<br />

oogonium are at first multinucleate. Nuclear<br />

division precedes fertilization, and meiosis<br />

occurs in the oogonium and antheridium<br />

(Sansome & Sansome, 1974). Fusion between<br />

two nuclei is delayed at least until the oospore<br />

wall is partly formed.<br />

The wall of the oospore of P. parasitica is very<br />

<strong>to</strong>ugh, and it is difficult <strong>to</strong> induce germination. In<br />

P. destruc<strong>to</strong>r and some other species, germination<br />

occurs by means of a germ tube but in P. tabacina<br />

zoospores have been described. It is probable<br />

that oospores overwinter in soil and give rise<br />

<strong>to</strong> infection in subsequent seasons. Although<br />

oospores of P. destruc<strong>to</strong>r have been germinated<br />

after 25 years, it has not proven possible <strong>to</strong> infect<br />

onions from such material. Possibly in this case<br />

the disease is carried over by means of systemic<br />

infection of volunteer onion bulbs (Smith et al.,<br />

1988).<br />

Peronospora parasitica and<br />

Arabidopsis thaliana<br />

The chance discovery of a P. parasitica infection in<br />

an Arabidopsis thaliana weed population in a<br />

Zurich garden showing haus<strong>to</strong>ria, sporangia<br />

and oospores (Koch & Slusarenko, 1990) opened<br />

up the possibility of using this genetically<br />

well-characterized ‘model plant’ <strong>to</strong> investigate<br />

plant pathogen interactions involving downy<br />

mildews. The interaction between Arabidopsis<br />

and Peronospora is governed by a gene-for-gene<br />

relationship, i.e. it is a form of major gene<br />

resistance based on specific recognition of a<br />

pathogen avirulence gene (avr) product by the<br />

product of a matching host resistance (R) gene<br />

(e.g. Botella et al., 1998). Molecular aspects of<br />

the Arabidopsis immune response <strong>to</strong> infections by


PERONOSPORALES<br />

119<br />

P. parasitica and other pathogens have been<br />

investigated in some detail. Infection of one<br />

leaf triggers a localized reaction, the hypersensitive<br />

response, leading <strong>to</strong> death of the plant<br />

cells in the vicinity of infection. Additionally,<br />

a systemic response is initiated, i.e. plant<br />

organs distal <strong>to</strong> the infected leaf become resistant<br />

against further attack. This phenomenon is<br />

called systemic acquired resistance and is active<br />

against attacks by the same as well as many<br />

other pathogens. It is triggered at the site of<br />

initial infection by various elici<strong>to</strong>r molecules of<br />

pathogen origin, e.g. fatty acids such as arachidonic<br />

acid, or by other substances. The signal is<br />

transmitted by signalling molecules such as<br />

salicylic acid (Law<strong>to</strong>n et al., 1995; Ton et al.,<br />

2002) which itself has no antimicrobial activity.<br />

Salicylic acid-independent signalling events are<br />

probably also involved (McDowell et al., 2000).<br />

Salicylic acid is produced at sites of infection,<br />

diffuses through the plant and interacts with a<br />

signalling chain, leading <strong>to</strong> the expression of a<br />

set of pathogenesis-related (PR) genes. A whole<br />

subset of PR genes involved in resistance <strong>to</strong><br />

P. parasitica (RPP genes) is now known (McDowell<br />

et al., 2000). The function of many PR genes is<br />

still obscure; those whose functions are known<br />

encode chitinases, b-1,3-glucanases, proteinases,<br />

peroxidases or enzymes involved in <strong>to</strong>xin<br />

biosynthesis (Kombrink & Somssich, 1997). By<br />

creating mutants of Arabidopsis or of crop plants<br />

which overexpress their own regula<strong>to</strong>ry genes<br />

or PR genes, or express introduced genes<br />

encoding elici<strong>to</strong>r molecules of pathogen origin,<br />

constitutive resistance against pathogen attack<br />

may be generated. This is considered <strong>to</strong> hold<br />

great potential for agriculture (Cao et al., 1998;<br />

Maleck et al., 2002).<br />

Control of Peronospora<br />

Downy mildew infections caused by Peronospora<br />

spp. are controlled mainly by fungicide applications.<br />

Metalaxyl is very effective against all<br />

downy mildews, but resistance has arisen in<br />

several species, and thus this fungicide is now<br />

applied in a cocktail with dithiocarbamates<br />

(Smith et al., 1988). Fosetyl Al is also now<br />

widely used as a foliar spray, root dip or soil<br />

amendment (Agrios, 2005).<br />

The breeding of cultivars with resistance<br />

against Peronospora spp. has been successful in<br />

certain crops, e.g. in lucerne (Medicago sativa)<br />

against P. trifoliorum (Stuteville, 1981). In <strong>to</strong>bacco<br />

plants attacked by P. tabacina, this strategy is a<br />

useful component of integrated control but is<br />

not sufficient on its own <strong>to</strong> afford complete<br />

control (Schiltz, 1981). In the <strong>to</strong>bacco P. tabacina<br />

system, a disease warning system is also in<br />

operation in Europe; subscribing <strong>to</strong>bacco<br />

growers are informed of the occurrence of the<br />

pathogen, so that preventative measures can be<br />

taken (Smith et al., 1988). This is profitable<br />

because <strong>to</strong>bacco is a high-value crop.<br />

Because downy mildews infect aerial plant<br />

parts and produce air-borne propagules in large<br />

numbers, crop sanitation measures are generally<br />

not very effective. However, in the case of<br />

P. destruc<strong>to</strong>r which overwinters systemically in<br />

volunteer onion bulbs, removal of volunteers is<br />

essential. In P. viciae on peas and beans, deep<br />

ploughing of the crop residue is important as the<br />

pathogen survives on infected haulms (Smith<br />

et al., 1988).<br />

5.4.2 Plasmopara (Peronosporaceae)<br />

Although downy mildews caused by species of<br />

Plasmopara are rarely serious in temperate<br />

climates, P. viticola is potentially a very destructive<br />

pathogen of the grapevine. The disease,<br />

which was endemic in North America and not<br />

particularly destructive on the local vines, was<br />

introduced in<strong>to</strong> France during the nineteenth<br />

century with disastrous results on the French<br />

vines which had never been exposed <strong>to</strong> the<br />

disease and were highly susceptible. Large (1940)<br />

has vividly recounted the moment when Alexis<br />

Millardet, walking past a heavily infected vineyard<br />

in 1882, noticed that vines close <strong>to</strong> the road<br />

appeared healthy and had been sprayed with a<br />

mixture of lime and copper sulphate <strong>to</strong> discourage<br />

passers-by from pilfering fruit. This led <strong>to</strong><br />

the discovery of Bordeaux mixture, one of the<br />

world’s first fungicides and still effective against<br />

P. viticola and other foliar pathogens belonging <strong>to</strong><br />

the Oomycota.<br />

Plasmopara nivea is occasionally reported in<br />

Britain on umbelliferous crops such as carrot


120 STRAMINIPILA: OOMYCOTA<br />

and parsnip, and it is also found on Aegopodium<br />

podagraria. Plasmopara pygmaea is found on<br />

yellowish patches on the leaves of Anemone<br />

nemorosa (Fig. 5.30b), whilst P. pusilla is similarly<br />

associated with Geranium pratense (Fig. 5.30a).<br />

The haus<strong>to</strong>ria of Plasmopara are knob-like,<br />

the sporangiophores are branched monopodially<br />

and the sporangia are hyaline (Fig. 5.30). Two<br />

types of sporangial germination have been<br />

reported. In P. pygmaea there are no zoospores<br />

but the entire sporangium detaches and later<br />

produces a germ-tube. In other species the<br />

sporangia germinate by means of zoospores<br />

which encyst and penetrate the host s<strong>to</strong>mata.<br />

Oospore germination in P. viticola is also by<br />

means of zoospores.<br />

Because the grapevine is such a highvalue<br />

crop, the fungicide market is lucrative.<br />

Bordeaux mixtures are still used <strong>to</strong>day, and<br />

similar fungicide applications <strong>to</strong> those described<br />

for Peronospora are made. Resistance <strong>to</strong> metalaxyl<br />

has been observed in P. viticola. Disease forecasting<br />

systems are being developed (Lafon & Bulit,<br />

1981; Smith et al., 1988). Breeding for resistant<br />

cultivars is being carried out, but because of the<br />

long generation times of the crop, this will be a<br />

prolonged effort.<br />

5.4.3 Bremia (Peronosporaceae)<br />

Bremia lactucae causes downy mildew of lettuce<br />

(Lactuca sativa) and strains of it can be found on<br />

36 genera of the Asteraceae including Sonchus<br />

and Senecio (Crute & Dixon, 1981). Crossinoculation<br />

experiments using sporangia from<br />

these hosts have failed <strong>to</strong> result in infection of<br />

lettuce and it seems that the fungus exists as a<br />

number of host-specific strains (formae speciales).<br />

Although wild species of Lactuca can carry strains<br />

capable of infecting lettuce, these hosts are not<br />

sufficiently common <strong>to</strong> provide a serious source<br />

of infection. The disease can be troublesome both<br />

in lettuce grown in the open and under frames,<br />

Fig 5.30 Plasmopara.<br />

(a) Sporangiophores<br />

of P. pusilla on<br />

Geranium pratense.<br />

(b) Sporangiophores<br />

of P. pygmaea on<br />

Anemone nemorosa.


PERONOSPORALES<br />

121<br />

and in market gardens there may be sufficient<br />

overlap in the growing of lettuce for the disease<br />

<strong>to</strong> be carried over from one sowing <strong>to</strong> the next.<br />

The damage <strong>to</strong> the crop caused by Bremia may<br />

not in itself be severe, but infected plants are<br />

prone <strong>to</strong> secondary infection by the more serious<br />

grey mould, Botrytis cinerea. Systemic infections<br />

can occur. The intercellular mycelium is coarse,<br />

and the haus<strong>to</strong>ria are sac-shaped, often several of<br />

them being present in each host cell (Fig. 5.31d).<br />

The sporangiophores emerge singly or in small<br />

groups through the s<strong>to</strong>mata and branch dicho<strong>to</strong>mously.<br />

The tip of each branch expands <strong>to</strong><br />

form a cup-shaped disc bearing short cylindrical<br />

Fig 5.31 Bremia lactucae from Senecio vulgaris. (a) Sporangiophore protruding through a s<strong>to</strong>ma. (b) Sporangiophore apex.<br />

(c) Sporangium germinating by means of a germ tube which has produced an appressorium at its apex. (d) Cells of epidermis and<br />

palisade mesophyll, showing intercellular mycelium and haus<strong>to</strong>ria. (a,c,d) <strong>to</strong> same scale.


122 STRAMINIPILA: OOMYCOTA<br />

sterigmata at the margin and occasionally in the<br />

centre, and from these the hyaline sporangia<br />

arise (Figs. 5.31a,b). Germination of the sporangia<br />

is usually by means of a germ tube which<br />

forms an appressorium <strong>to</strong> penetrate epidermal<br />

cells (Fig. 5.31c), or it enters through a s<strong>to</strong>ma.<br />

Zoospore formation has been reported but<br />

not confirmed. Sexual reproduction is usually<br />

heterothallic, although homothallic strains also<br />

exist. The oospores are formed in leaf tissue and<br />

remain viable for 12 months (Michelmore &<br />

Ingram, 1980; Morgan, 1983).<br />

Chemical control of B. lactucae on lettuce is<br />

certainly possible although not necessarily desirable;<br />

hence, intensive efforts for major gene<br />

resistance breeding have been made. Integrated<br />

control based on resistant cultivars and fungicide<br />

applications using metalaxyl and dithiocarbamates<br />

is successful (Crute, 1984). However,<br />

resistance against metalaxyl arose in Britain as<br />

early as 1983. Fosetyl Al is not as effective as<br />

metalaxyl (Smith et al., 1988).<br />

5.4.4 Albugo (Albuginaceae)<br />

This family has only a single genus, Albugo, with<br />

about 40 50 species of biotrophic parasites of<br />

flowering plants which cause diseases known as<br />

white blisters or white rusts. The commonest<br />

British species is A. candida causing white blisters<br />

of crucifers such as cabbage, turnip, swede,<br />

horseradish, etc. (Plate 2h). It is particularly<br />

frequent on shepherd’s purse (Capsella bursapas<strong>to</strong>ris).<br />

There is some degree of physiological<br />

specialization in the races of this fungus on<br />

different host genera. Albugo candida can infect<br />

Arabidopsis thaliana, and the host defence<br />

response is governed by resistance genes involved<br />

in the recognition of the pathogen (Holub et al.,<br />

1995). The principle is similar <strong>to</strong>, although not as<br />

well researched as, the Arabidopsis Peronospora<br />

interaction described earlier (p. 116). It is also<br />

now possible <strong>to</strong> establish callus cultures of mustard<br />

plants (Brassica juncea) containing balanced<br />

infections of A. candida (Nath et al., 2001). This<br />

experimental system should facilitate studies<br />

of the physiology of host pathogen interactions.<br />

A less common species is A. tragopogonis, causing<br />

white blisters of salsify (Tragopogon porrifolius),<br />

goatbeard (T. pratensis) and Senecio squalidus.<br />

In A. candida on shepherd’s purse, diseased<br />

plants may be detected by the dis<strong>to</strong>rted stems<br />

and the shining white raised blisters on the stem,<br />

leaves and pods before the host epidermis is<br />

ruptured (Plate 2h). Later, when the epidermis<br />

has burst open, a white powdery pustule is<br />

visible. The dis<strong>to</strong>rtion is possibly associated with<br />

altered auxin levels. The host plant may be<br />

infected simultaneously with Peronospora parasitica,<br />

but the two fungi are easily distinguishable<br />

microscopically both in the structure of the<br />

sporangiophores and by their different haus<strong>to</strong>ria.<br />

In Albugo, the mycelium in the host tissues<br />

is intercellular with only small spherical haus<strong>to</strong>ria<br />

(Fig. 5.32) which contrast sharply with the<br />

coarsely lobed haus<strong>to</strong>ria of P. parasitica. The fine<br />

structure of A. candida haus<strong>to</strong>ria has been<br />

described by Coffey (1975) and Soylu et al.<br />

(2003). They are spherical or somewhat flattened<br />

and about 4 mm in diameter, connected <strong>to</strong> the<br />

intercellular mycelium by a narrow stalk about<br />

0.5 mm wide. Inside the plasma membrane of the<br />

haus<strong>to</strong>rium, lomasomes, i.e. tubules and vesicles<br />

apparently formed by invagination of the plasma<br />

membrane, are more numerous than in the<br />

intercellular hyphae. The cy<strong>to</strong>plasm of the haus<strong>to</strong>rial<br />

head is densely packed with mi<strong>to</strong>chondria,<br />

ribosomes, endoplasmic reticulum and occasional<br />

lipid droplets, but nuclei have not been<br />

observed. Since nuclei of Albugo are about 2.5 mm<br />

in diameter, they may be unable <strong>to</strong> traverse the<br />

constriction which links the haus<strong>to</strong>rium <strong>to</strong> the<br />

intercellular hypha. Nuclei may (e.g. Peronospora<br />

pisi) or may not be present in the haus<strong>to</strong>ria of<br />

other Oomycota. The base of the haus<strong>to</strong>rium of<br />

A. candida is surrounded by a collar-like sheath<br />

which is an extension of the host cell wall, but<br />

this wall does not normally extend <strong>to</strong> the main<br />

body of the haus<strong>to</strong>rium. Between the haus<strong>to</strong>rium<br />

and the host plasma membrane is an encapsulation.<br />

Host cy<strong>to</strong>plasm reacts <strong>to</strong> infection by an<br />

increase in the number of ribosomes and Golgi<br />

complexes. In the vicinity of the haus<strong>to</strong>rium<br />

the host cy<strong>to</strong>plasm contains numerous vesicular<br />

and tubular elements not found in uninfected<br />

cells. These structures have been interpreted


PERONOSPORALES<br />

123<br />

Fig 5.32 Albugo candida on Capsella bursa-pas<strong>to</strong>ris. (a) Mycelium, sporangiophores and chains of sporangia formed beneath the<br />

ruptured epidermis (right). (b) Germination of sporangia showing the release of eight biflagellate zoospores.The stages illustrated<br />

<strong>to</strong>ok place within 2 min. (c) Haus<strong>to</strong>ria.<br />

as evidence of secre<strong>to</strong>ry processes induced in the<br />

host cell by the presence of the pathogen.<br />

The intercellular mycelium aggregates<br />

beneath the host epidermis <strong>to</strong> form a palisade<br />

of cylindrical or skittle-shaped sporangiophores<br />

which give rise <strong>to</strong> chains of spherical sporangia<br />

in basipetal succession i.e. new sporangia<br />

are formed at the base of the chain. The pressure<br />

of the developing chains of sporangia raises<br />

the host epidermis and finally ruptures it.


124 STRAMINIPILA: OOMYCOTA<br />

The sporangia are then visible externally as a<br />

white powdery mass dispersed by the wind.<br />

Sporangia reaching a suitable host leaf will<br />

germinate within a few hours in films of water<br />

<strong>to</strong> form biflagellate zoospores of the principal<br />

type, about eight per sporangium (Fig. 5.32b).<br />

After swimming for a time, a zoospore encysts<br />

and then forms a germ tube which penetrates<br />

the host epidermis. The asexual disease cycle<br />

may be completed within 10 days. Infections may<br />

be localized or systemic. Gametangia are formed<br />

in the intercellular spaces of infected stems and<br />

leaves. Both the antheridium and the oogonium<br />

are multinucleate at their inception, and during<br />

development two further nuclear divisions occur<br />

so that the oogonium may contain over 200<br />

nuclei. However, there is only one functional<br />

male and one functional female nucleus. In the<br />

oogonium all the nuclei except one migrate <strong>to</strong><br />

the periphery and are included in the periplasm.<br />

Following nuclear fusion a thin membrane first<br />

develops around the oospore. Division of the<br />

zygote nucleus takes place and is repeated, so<br />

that at maturity the oospore may contain as<br />

many as 32 diploid nuclei. Sansome and<br />

Sansome (1974) reported that meiosis occurs<br />

within the gametangia. They also suggested<br />

that A. candida is heterothallic. The high incidence<br />

of oospores of Albugo in Capsella stems<br />

simultaneously infected with Peronospora<br />

parasitica may result from some stimulus<br />

<strong>to</strong>wards self-fertilization in Albugo produced by<br />

Peronospora, a situation analogous <strong>to</strong> the<br />

Trichoderma-induced sexual reproduction in<br />

heterothallic species of Phy<strong>to</strong>phthora (see p. 95).<br />

The mature oospore is surrounded by a brown<br />

exospore, thrown in<strong>to</strong> warty folds (Fig. 5.33a).<br />

Germination of the oospores takes place only<br />

after a resting period of several months. Under<br />

suitable conditions the outer wall of the oospore<br />

bursts and the endospore is extruded as<br />

a thin, spherical vesicle, which may be sessile<br />

or formed at the end of a wide cylindrical tube.<br />

Within the thin vesicle 40 60 zoospores are<br />

differentiated and are released on its breakdown<br />

(Figs. 5.33b,c).<br />

The cy<strong>to</strong>logy of oospore development in<br />

some other species of Albugo differs from that<br />

of A. candida. InA. bliti, a pathogen of Portulaca<br />

in North America and Europe, the oogonia and<br />

antheridia are also multinucleate and two<br />

nuclear divisions take place during their development.<br />

Numerous male nuclei fuse with<br />

numerous female nuclei and the fusion nuclei<br />

Fig 5.33 Albugo candida oospores. (a) Oogonium and oospore<br />

from Capsella leaf. (b,c) Two methods of oospore germination<br />

(after Vanterpool,1959).


SCLEROSPORACEAE<br />

125<br />

pass the winter without further change. In<br />

A. tragopogonis, a multinucleate oospore develops<br />

and again there are two nuclear divisions<br />

involved in the development of the oogonium<br />

and antheridium, but finally there is a single<br />

nuclear fusion between one male and one<br />

female nucleus. This fusion nucleus undergoes<br />

repeated divisions so that the overwintering<br />

oospore is multinucleate.<br />

Albugo candida alone or in combination with<br />

co-infecting Peronospora parasitica can occasionally<br />

cause significant crop losses in cabbage<br />

cultivation. <strong>Fungi</strong>cide treatment is possible,<br />

with copper-based or dithiocarbamate-type fungicides<br />

commonly used (Smith et al., 1988).<br />

5.5 Sclerosporaceae<br />

This family comprises the downy mildews of<br />

grasses and cereals. Although it is well defined as<br />

a biological group, its phylogenetic position is<br />

unclear, recent ribosomal DNA-based studies<br />

placing its members among the Peronosporales<br />

(Riethmüller et al., 2002). For reasons of their<br />

distinctly different biological features, we<br />

consider them briefly here. The principal<br />

genera are Sclerospora, with sporangia capable of<br />

germinating by releasing zoospores, and<br />

Peronosclerospora, whose sporangia show direct<br />

germination by germ tubes and are thus,<br />

functionally speaking, ‘conidia’. Sporangia or<br />

conidia are produced on repeatedly branching<br />

aerial structures which resemble those of<br />

Peronospora spp. In Peronosclerospora, the conidiophores<br />

project through s<strong>to</strong>mata of the host and<br />

branch at their apices <strong>to</strong> produce up <strong>to</strong> 20 fingerlike<br />

tapering extensions which expand <strong>to</strong> form<br />

conidia (Figs. 5.34a c). The conidia are oval<br />

and hyaline. Unlike those of other Oomycota,<br />

conidia of Sclerosporaceae are projected actively<br />

by a sudden rounding-off of the conidiophore<br />

tip and conidial base, and this is visible as a<br />

Fig 5.34 Peronosclerospora sorghi.<br />

(a) Immature conidiophore showing<br />

conidium initials. (b) Mature conidiophore<br />

from which two conidia have become<br />

detached. (c) Old conidiophore; all conidia<br />

have become detached. (d) Discharged<br />

conidia. Note the small basal projection.<br />

Drawn from material kindly provided<br />

by K. Mathur.


126 STRAMINIPILA: OOMYCOTA<br />

Fig 5.35 Oospore of Peronosclerospora sorghi.Notethe<br />

thickened oogonial wall (arrow), within which the spherical<br />

oospore with its wall and ooplast is clearly visible.<br />

small projection at the base of discharged<br />

conidia (Fig. 5.34d). Oospores of Sclerosporaceae<br />

are distinctive in being surrounded by a thickened<br />

oogonial wall (Fig. 5.35), and this feature<br />

may enhance the longevity of the oospore. The<br />

most important species are Sclerospora graminicola<br />

infecting pearl millet (Pennisetum americanum),<br />

and Peronosclerospora sorghi pathogenic on<br />

sorghum and maize. Because of their similar<br />

biological features and great economic importance,<br />

these two species are often considered<br />

<strong>to</strong>gether. Thorough reviews have been written<br />

by R. J. Williams (1984) and Jeger et al. (1998).<br />

Downy mildews of grasses cause serious<br />

crop losses especially in dry subtropical and<br />

tropical zones in Africa, their putative centre of<br />

evolution, as well as Asia and, <strong>to</strong> a lesser extent,<br />

North and South America. The thick-walled<br />

oospores can survive on plant debris and in the<br />

soil for up <strong>to</strong> 10 years, and infections are usually<br />

initiated from oospores which germinate<br />

directly by means of a germ tube. The plant<br />

root may be the initial route of entry, although<br />

both S. graminicola and P. sorghi may also become<br />

seed-borne. Later infections are through the<br />

shoot surface, either by direct penetration of<br />

the epidermis by means of appressoria, or<br />

through s<strong>to</strong>mata. Infections of host plants are<br />

obligately biotrophic and can become systemic<br />

if they reach the apical meristem. Sporangia<br />

or conidia are formed only on freshly infected<br />

living host tissues under moist conditions, and<br />

infections are therefore polycyclic only when<br />

sufficient moisture is available. In dry regions,<br />

infections may be carried exclusively by<br />

oospores, confining the pathogen <strong>to</strong> one disease<br />

cycle per growing season. Oospore production<br />

is buffered against environmental extremes by<br />

taking place within the tissue of aerial host<br />

organs. Like sporangia or conidia, oospores can<br />

be blown about by wind.<br />

Control of downy mildew of grasses is<br />

difficult. Metalaxyl gives good control both as a<br />

seed dressing and as a foliar spray but may<br />

not always be available. Numerous cultivars of<br />

sorghum and pearl millet show resistance<br />

against downy mildews, but this is usually<br />

based on one or a few major genes and can<br />

therefore be overcome by the pathogens if single<br />

cultivars are grown in large coherent areas.<br />

On small-scale farms, it may be possible <strong>to</strong> remove<br />

individual infected plants prior <strong>to</strong> the onset of<br />

sporulation (Gilijamse et al., 1997).


6<br />

Chytridiomycota<br />

6.1 <strong>Introduction</strong><br />

The phylum Chytridiomycota comprises over<br />

900 species in five orders (D. J. S. Barr, 2001; Kirk<br />

et al., 2001). <strong>Fungi</strong> included here are colloquially<br />

called ‘chytrids’. Most chytrids grow aerobically<br />

in soil, mud or water and reproduce by zoospores<br />

with a single posterior flagellum of the whiplash<br />

type, although the zoospores of some members of<br />

the Neocallimastigales are multiflagellate. Some<br />

species inhabit estuaries and others the sea.<br />

Sparrow (1960) has given an extensive account<br />

of aquatic forms, Karling (1977) a compendium of<br />

illustrations, and Powell (1993) has provided<br />

examples of the importance of the group. Many<br />

members are saprotrophs, utilizing cellulose,<br />

chitin, keratin, etc., from decaying plant and<br />

animal debris in soil and mud, whilst species of<br />

Caulochytrium grow as mycoparasites on the<br />

mycelium and conidia of terrestrial fungi (Voos,<br />

1969). Saprotrophs can be obtained in crude<br />

culture by floating baits such as cellophane,<br />

hair, shrimp exoskele<strong>to</strong>n, boiled grass leaves<br />

and pollen on the surface of water overlying<br />

samples of soil, mud or pieces of aquatic plant<br />

material (Sparrow, 1960; Stevens, 1974;<br />

Willoughby, 2001). From such crude material,<br />

pure cultures may be prepared by streaking<br />

or pipetting zoospores on<strong>to</strong> agar containing<br />

suitable nutrients and antibiotics <strong>to</strong> limit contamination<br />

from bacteria. The growth and<br />

appearance of chytrids in pure culture is variable<br />

and often differs significantly from their natural<br />

habit. This has led <strong>to</strong> problems in classification<br />

systems based on thallus morphology (Barr, 1990,<br />

2001). The availability of cultures has, however,<br />

facilitated studies on chytrid nutrition and<br />

physiology (Gleason, 1976).<br />

Some chytrids are biotrophic parasites of<br />

filamen<strong>to</strong>us algae and dia<strong>to</strong>ms and may severely<br />

deplete the population of freshwater phy<strong>to</strong>plank<strong>to</strong>n<br />

(see p. 139). Two-membered axenic<br />

cultures of dia<strong>to</strong>m host and parasite have been<br />

prepared, making possible detailed ultrastructural<br />

studies of comparative morphology, zoospores,<br />

infection processes and reproduction.<br />

Other chytrids such as species of Synchytrium<br />

and Olpidium are biotrophic parasites of<br />

vascular plants. Synchytrium endobioticum is the<br />

agent of the potentially serious black wart<br />

disease of pota<strong>to</strong>. Olpidium brassicae, common in<br />

the roots of many plants, is relatively harmless,<br />

but its zoospores are vec<strong>to</strong>rs of viruses such<br />

as that causing big vein disease of lettuce.<br />

Coelomomyces spp. are pathogens of freshwater<br />

invertebrates including copepods and the larvae<br />

of mosqui<strong>to</strong>es. The possibility of using them<br />

in the biological control of mosqui<strong>to</strong>es has<br />

been explored. The most unusual group are the<br />

Neocallimastigales, which grow in the guts of<br />

herbivorous mammals, are obligately anaerobic<br />

and subsist on ingested herbage.<br />

The cell walls of some chytrids have been<br />

examined microchemically by X-ray diffraction<br />

and other techniques. Chitin has been detected<br />

in many species (Bartnicki-Garcia, 1968, 1987),<br />

and in Gonapodya cellulose is also present<br />

(Fuller & Clay, 1993). The composition of the<br />

wall is of interest because chitin, a polymer of


128 CHYTRIDIOMYCOTA<br />

N-acetylglucosamine, is also found in the walls<br />

of other Eumycota (i.e. Zygomycota, Ascomycota<br />

and Basidiomycota), whilst the cell walls of<br />

members of the Oomycota contain cellulose.<br />

Cellulose and chitin occur <strong>to</strong>gether in the walls<br />

of species of Hyphochytrium and Rhizidiomyces,<br />

members of the Hyphochytriomycota (Fuller,<br />

2001; see Section 4.3).<br />

The form of the thallus in the<br />

Chytridiomycota is varied. In biotrophic species<br />

such as Olpidium and Synchytrium, where the<br />

whole thallus is contained within the host cell,<br />

there is no differentiation in<strong>to</strong> a vegetative<br />

and a reproductive part. At maturity the entire<br />

structure, except for the wall which surrounds it,<br />

is converted in<strong>to</strong> reproductive units, i.e. zoospores,<br />

gametes or resting sporangia. Such thalli<br />

are termed holocarpic (Fig. 6.1). More usually, the<br />

thallus is differentiated in<strong>to</strong> organs of reproduction<br />

(sporangia and resting sporangia) arising<br />

from a vegetative part which often consists of<br />

rhizoids. These serve in the exploitation of the<br />

substratum and the assimilation of nutrients.<br />

Thalli of this type are eucarpic. Eucarpic thalli<br />

may have one or several sporangia and are then<br />

termed monocentric or polycentric, respectively<br />

(Fig. 6.1). In some species there are both monocentric<br />

and polycentric thalli, so that these terms<br />

have descriptive rather than taxonomic significance.<br />

A further distinction has been made,<br />

especially in monocentric forms, between those<br />

in which only the rhizoids are inside the host<br />

cell whilst the sporangium is external (epibiotic),<br />

in contrast with the endobiotic condition in<br />

which the entire thallus is inside the host cell<br />

(Fig. 6.1). In monocentric thalli, the rhizoids<br />

usually radiate from a single position on the<br />

sporangium wall, but in polycentric forms a<br />

more extensive, branched rhizoidal system, the<br />

rhizomycelium, develops.<br />

The zoosporangium is generally a spherical<br />

or pear-shaped sac bearing one or more discharge<br />

tubes or exit papillae. The method of<br />

zoospore release has been used in classification.<br />

Fig 6.1 Types of thallus structure in the<br />

Chytridiales, diagrammatic and not <strong>to</strong> scale.


INTRODUCTION<br />

129<br />

In the inoperculate chytrids such as Olpidium,<br />

Diplophlyctis and Cladochytrium, the sporangium<br />

forms a discharge tube which penetrates <strong>to</strong> the<br />

exterior of the host cell and its tip becomes<br />

gelatinous and dissolves away. In operculate<br />

chytrids such as Chytridium and Nowakowskiella,<br />

the tip of the discharge tube breaks open at a<br />

special line of weakness and is seen as a special<br />

cap or operculum after discharge (see Fig. 6.4b).<br />

6.1.1 The zoospore<br />

The number of zoospores formed inside zoosporangia<br />

of chytrids varies with the size of the<br />

spore and sporangium. Although the zoospore<br />

size is roughly constant for a given species, the<br />

size of the sporangium may be very variable. In<br />

Rhizophlyctis rosea, tiny sporangia containing only<br />

one or two zoospores have been reported from<br />

culture media deficient in carbohydrate, whereas<br />

on cellulose-rich media large sporangia containing<br />

many hundred spores are formed. The release<br />

of zoospores is brought about by internal<br />

pressure which causes the exit papillae <strong>to</strong> burst<br />

open. In studies of the fine structure of mature<br />

sporangia of R. rosea and Nowakowskiella profusa<br />

(Chambers & Willoughby, 1964; Chambers et al.,<br />

1967), it has been shown that the single<br />

flagellum is coiled round the zoospore like a<br />

watch spring. The zoospores are separated by a<br />

matrix of spongy material which may absorb<br />

water and swell rapidly at the final stages<br />

of sporangial maturation. When the internal<br />

pressure has been relieved by the ejection of<br />

some zoospores, those remaining inside the<br />

sporangium escape by swimming or wriggling<br />

through the exit tube. In some species the spores<br />

are discharged in a mass which later separates<br />

in<strong>to</strong> single zoospores, but in others the zoospores<br />

make their escape individually.<br />

The form of the zoospore is similar in all<br />

chytrids (with the exception of the multiflagellate<br />

members of the Neocallimastigales). There<br />

is a spherical or ellipsoidal body which in some<br />

forms is capable of plastic changes in shape, and<br />

a long trailing flagellum. When swimming, the<br />

zoospores show characteristic jerky or ‘hopping’<br />

movements; additionally, abrupt changes in<br />

direction are sometimes made. The internal<br />

structure of the zoospore as revealed by light<br />

and electron microscopy is variable, but characteristic<br />

of particular genera (Lange & Olson,<br />

1979). In view of the plasticity in morphology of<br />

the thallus under different growth conditions,<br />

zoospore ultrastructure is regarded as a more<br />

satisfac<strong>to</strong>ry basis of classification (D. J. S. Barr,<br />

1990, 2001). Two features are of taxonomic<br />

importance, the flagellar apparatus and an assemblage<br />

of organelles termed the microbody lipid<br />

globule complex (MLC) (D. J. S. Barr, 2001).<br />

The flagellar apparatus<br />

The whiplash flagellum resembles that of other<br />

eukaryotes, with a smooth membrane enclosing<br />

a cylindrical shaft, the axoneme, made up<br />

internally of nine doublet pairs of microtubules<br />

surrounding two central microtubules. As shown<br />

in Fig. 6.2, the base of the axoneme comprises<br />

three regions, the flagellum proper, the transitional<br />

zone and the kine<strong>to</strong>some. The function<br />

of the kine<strong>to</strong>some is <strong>to</strong> generate the flagellum.<br />

An interesting feature found in several species<br />

is a second kine<strong>to</strong>some or the remainder of<br />

one, the dormant kine<strong>to</strong>some. Its presence<br />

has led <strong>to</strong> the suggestion that the ances<strong>to</strong>rs<br />

of the Chytridiomycota may have had biflagellate<br />

zoospores, the second flagellum having been<br />

lost in the course of evolution (Olson & Fuller,<br />

1968).<br />

In section, the kine<strong>to</strong>some resembles a cartwheel<br />

(Fig. 6.2f), because <strong>to</strong> each of the nine<br />

outer microtubule doublets seen in the flagellum<br />

proper, a third microtubule is attached.<br />

This is called the C-tubule; in the doublets,<br />

that tubule with extended dynein arms is the<br />

A-tubule, and its partner is labelled B. These<br />

flagellar microtubules radiate as kine<strong>to</strong>some<br />

props in<strong>to</strong> the zoospore, perhaps providing<br />

structural support and anchorage of the flagellum<br />

(D. J. S. Barr, 2001). Microtubules may also be<br />

attached laterally <strong>to</strong> the kine<strong>to</strong>some, contributing<br />

<strong>to</strong> the flagellar root system (Figs. 6.2c, 6.19).<br />

In the innermost (proximal) part of the transitional<br />

zone, the nine microtubule triplets of<br />

the kine<strong>to</strong>some are converted in<strong>to</strong> the doublets<br />

of the flagellum proper; concentric fibres,<br />

possibly arranged helically, surround the nine<br />

doublet pairs. Also within the transitional zone,


130 CHY TRIDIOMYCOTA<br />

Fig 6.2 Flagellar apparatus typical of zoospores of Chytridiomycota. (a) Median longitudinal section of the junction of the flagellum<br />

with the body of the zoospore.The labels indicate the flagellum proper (F), transitional zone (TZ), kine<strong>to</strong>some (K), electron-dense<br />

region (ED), concentric fibres (CF), transitional fibres (TF), kine<strong>to</strong>some props (KP), terminal plate (TP), kine<strong>to</strong>some (K) showing a<br />

cartwheel-like organization (Cw), dormant kine<strong>to</strong>some (DK), fibrillar material (Fi) found in some taxa, and microtubular roots (Mt)<br />

extending from the side or end of the kine<strong>to</strong>some in<strong>to</strong> the body of the zoospore. (b) Transverse section near the terminal plate<br />

showing nine kine<strong>to</strong>some props extending from doublet microtubules <strong>to</strong> the cell membrane. (c) Transverse section in the lower<br />

part of the transition zone showing concentric and transitional fibres. (d) Transverse section of the flagellum proper showing two<br />

central microtubules and nine peripheral doublet microtubules enclosed in the flagellar membrane (FM). (e) Schematic drawing of<br />

the flagellum proper in transverse section.The arrowed line 0° 180° shows an imaginary plane which coincides with the plane of<br />

undulation of the flagellum, passing through doublet pair1and between the central microtubules and doublet pairs 5 and 6.<br />

The convention used in labelling the outer doublet pairs of microtubules is shown: the microtubule with dynein arms (d) is the<br />

A microtubule and its partner is the B microtubule. (f) Kine<strong>to</strong>some in transverse section showing the triplet arrangement of the<br />

peripheral microtubules by the addition of a third microtubule (C). Redrawn from Barr and De¤saulniers (1988) by copyright<br />

permission of the National Research Council of Canada, Barr (1992). ßThe Mycological Society of America, and D. J. S. Barr (2001)<br />

with kind permission of Springer Science and Business Media.<br />

the two central microtubules arise near a<br />

terminal plate. The structure of the flagellum<br />

and kine<strong>to</strong>some in transverse section is shown in<br />

Figs. 6.2e and f (Barr & Désaulniers, 1988).<br />

The microbody lipid complex<br />

The MLC (Fig. 6.3) is made up of a microbody<br />

which is often closely appressed <strong>to</strong> a large lipid<br />

globule and <strong>to</strong> simple membrane cisternae or<br />

a tubular membrane system, the rumposome.<br />

This is defined as a cisterna in which there is an<br />

area with hexagonally arranged, honeycomb-like<br />

pores called fenestrae (Fuller, 1976; Powell &<br />

Roychoudhury, 1992). The rumposome may<br />

be involved in signal transduction from the<br />

plasma membrane <strong>to</strong> the flagellum because it<br />

is known that this organelle sequesters calcium.<br />

Regulation of external calcium concentrations<br />

has an effect on the symmetry of flagellar beat<br />

and hence on the direction of zoospore movement<br />

(Powell, 1983).<br />

There are several distinct types of MLC (Powell<br />

& Roychoudhury, 1992) and Fig. 6.3 illustrates<br />

diagrammatically just one of them, that<br />

described for Rhizophlyctis harderi. In this species,<br />

the MLC includes several (3 5) lipid globules.


INTRODUCTION<br />

131<br />

Most zoospores are uninucleate. The nucleus<br />

is surrounded in many cases (but not all) by a<br />

nuclear cap of uneven thickness. The nuclear cap<br />

is especially prominent in zoospores of members<br />

of Blas<strong>to</strong>cladiales such as Allomyces and<br />

Blas<strong>to</strong>cladiella (Fig. 6.19). It is rich in RNA and<br />

protein and also contains ribosomes.<br />

Fig 6.3 Schematic diagram of the microbody lipid complex<br />

of the zoospore of Rhizophlyctis harderi as seen in a longitudinal<br />

section through the base of the zoospore and flagellum.The<br />

following organelles are drawn: mi<strong>to</strong>chondrion (Mc), simple<br />

cisterna (C), lipid globule (L), microbody (Mi), flagellum (F) and<br />

rumposome (R). Redrawn from Powell and Roychoudhury<br />

(1992), by copyright permission of the National Research<br />

Council of Canada.<br />

Those at the anterior of the cell are embedded<br />

in an aggregation of ribosomes. The surfaces<br />

of lipid globules close <strong>to</strong> the plasma membrane<br />

are partially covered by one <strong>to</strong> several simple<br />

cisternae, sometimes with irregularly scattered<br />

pores. Towards the centre of the cell the<br />

lipid bodies are clasped by cup-shaped microbodies.<br />

At the posterior of the zoospore near<br />

the kine<strong>to</strong>some, 1 3 smaller lipid globules<br />

are partially covered by a rumposome, linked<br />

<strong>to</strong> the plasma membrane by short bridges and<br />

<strong>to</strong> the kine<strong>to</strong>some by a microtubule root.<br />

Other features<br />

Patches of glycogen are located in the peripheral<br />

cy<strong>to</strong>plasm of the zoospore and it is likely<br />

that these and the lipid globules represent<br />

sources of energy used in respiration and<br />

propulsion. Mi<strong>to</strong>chondria tend <strong>to</strong> be concentrated<br />

in the posterior of the zoospore close<br />

<strong>to</strong> the kine<strong>to</strong>some; in Allomyces and Blas<strong>to</strong>cladiella<br />

(Blas<strong>to</strong>cladiales), the base of the flagellum<br />

passes through the perforation of a single large<br />

mi<strong>to</strong>chondrion (see Fig. 6.19).<br />

6.1.2 Zoospore encystment and<br />

germination<br />

The period of zoospore movement varies. Some<br />

flagellate zoospores seem <strong>to</strong> be incapable of<br />

active swimming and amoeboid crawling may<br />

take place instead, or swimming may last for<br />

only a few minutes. In other spores, motility<br />

may be prolonged for several hours. Prior <strong>to</strong><br />

germination, the zoospore comes <strong>to</strong> rest and<br />

encysts. The flagellum may contract, it may be<br />

completely withdrawn or it may be cast off,<br />

but the precise details are often difficult <strong>to</strong><br />

follow. The subsequent behaviour also differs<br />

in different species. In holocarpic parasites the<br />

zoospore encysts on the host surface and<br />

the cy<strong>to</strong>plasmic contents of the zoospore are<br />

injected in<strong>to</strong> the host cell. In many monocentric<br />

chytrids rhizoids develop from one point on<br />

the zoospore cyst and the cyst itself enlarges <strong>to</strong><br />

form the zoosporangium, but there are variants<br />

of this type of development in which the cyst<br />

enlarges in<strong>to</strong> a prosporangium from which the<br />

zoosporangium later develops. In the polycentric<br />

types, the zoospore on germination may form<br />

a limited rhizomycelium on which a swollen<br />

cell arises, giving off further branches of rhizomycelium.<br />

Germination may be from a single<br />

point on the wall of the zoospore cyst (monopolar<br />

germination) or from two points, enabling<br />

growth <strong>to</strong> take place in two directions (bipolar<br />

germination). The mode of germination is an<br />

important character in distinguishing, for<br />

example, the Chytridiales (monopolar) from<br />

the Blas<strong>to</strong>cladiales (bipolar).<br />

6.1.3 Life cycles of the Chytridiomycota<br />

Most chytrids have haploid zoospores and thalli<br />

but some Blas<strong>to</strong>cladiales show an alternation<br />

of haploid (game<strong>to</strong>thallic) and diploid (sporothallic)<br />

generations. Apart from differences in


132 CHY TRIDIOMYCOTA<br />

the reproductive organs, the morphology of the<br />

two types of thallus is very similar, a phenomenon<br />

known as isomorphic alternation of<br />

generations.<br />

Sexual reproduction, i.e. a life cycle which<br />

includes nuclear fusion and meiosis, may occur<br />

in several different ways (e.g. Figs. 6.6 and 6.22).<br />

In some chytrids it is by game<strong>to</strong>gamy, the fusion<br />

of gametes which are posteriorly uniflagellate.<br />

Isogamous conjugation occurs if there is no<br />

morphological distinction between the two<br />

fusing partners, but in some Blas<strong>to</strong>cladiales (e.g.<br />

Allomyces) anisogamy takes place by fusion<br />

between a smaller, more actively motile male<br />

gamete with a larger, sluggish female gamete.<br />

Oogamy, fusion between an actively motile<br />

male gamete and a much larger, non-flagellate,<br />

immobile globose egg, is characteristic of<br />

Monoblepharidales. Soma<strong>to</strong>gamy, the fusion<br />

of undifferentiated hyphae or rhizoids, has<br />

been well documented in cultures of the freshwater<br />

fungus Chytriomyces hyalinus by Moore<br />

and Miller (1973) and Miller and Dylewski<br />

(1981, 1987). As shown in Fig. 6.4, zoospores of<br />

C. hyalinus are released from the zoosporangium<br />

by the opening of a lid-like operculum. They<br />

germinate <strong>to</strong> form uninucleate rhizoidal thalli<br />

(contribu<strong>to</strong>ry thalli) and the tips of the rhizoids<br />

from adjacent thalli, which are apparently not<br />

genetically distinct from each other, may fuse<br />

(Fig. 6.4c). At the point of fusion an incipient<br />

resting body develops (Fig. 6.4d) and swells<br />

while cy<strong>to</strong>plasm and a nucleus migrate in<strong>to</strong> it<br />

from each contribu<strong>to</strong>ry thallus. Nuclear fusion<br />

occurs in the resting body <strong>to</strong> form a diploid<br />

zygote nucleus. The resting body continues <strong>to</strong><br />

enlarge and develops a thick wall. This type<br />

of sexual reproduction by soma<strong>to</strong>gamous conjugation<br />

probably occurs in several genera of<br />

inoperculate and operculate chytrids (Moore &<br />

Miller, 1973).<br />

Fusion of gametangia (gametangiogametangiogamy)<br />

has been reported by Doggett<br />

and Porter (1996) for Zygorhizidium plank<strong>to</strong>nicum,<br />

a parasite of the dia<strong>to</strong>m Synedra. This species<br />

reproduces asexually by epibiotic zoosporangia.<br />

Germinating zoospores develop either new<br />

zoosporangial thalli or gametangial thalli of<br />

two sizes with globose uninucleate gametangia.<br />

Fig 6.4 Chytriomyces hyalinus soma<strong>to</strong>gamy. (a,b) Epibiotic<br />

fruiting thallus seated on a pollen grain in<strong>to</strong> which rhizoids<br />

have penetrated. In (a) the zoosporangium, containing<br />

numerous zoospores, is seen shortly before discharge with a<br />

bulging operculum (o). In (b) the operculum has lifted off and<br />

the zoospores are escaping. (c e) Stages in soma<strong>to</strong>gamy.<br />

(c) Rhizoids from two uninucleate contribu<strong>to</strong>ry thalli (c) have<br />

undergone anas<strong>to</strong>mosis (arrow). (d) Cy<strong>to</strong>plasm and a nucleus<br />

from each contribu<strong>to</strong>ry thallus have migrated <strong>to</strong>wards the<br />

point of anas<strong>to</strong>mosis, where the thallus swells <strong>to</strong> form a<br />

globose incipient resting body (i) which is binucleate and<br />

packed with cy<strong>to</strong>plasm, leaving the contribu<strong>to</strong>ry thalli empty.<br />

(e) The two nuclei in the incipient resting body have fused.<br />

After C.E. Miller and Dylewski (1981).<br />

Conjugation occurs when a conjugation tube<br />

grows from the smaller donor <strong>to</strong> the larger<br />

recipient gametangium (Fig. 6.5a). Following<br />

nuclear fusion, the larger gametangium develops<br />

a thick wall and functions as a diploid resting<br />

spore. After a period of maturation the<br />

resting spore acts as a prosporangium, giving<br />

rise <strong>to</strong> a thin-walled meiosporangium. Meiosis,<br />

as evidenced by the presence of synap<strong>to</strong>nemal<br />

complexes, occurs here, followed by mi<strong>to</strong>sis<br />

and cy<strong>to</strong>plasmic cleavage <strong>to</strong> form zoospores<br />

(Fig. 6.5b). A variant of this form of sexual<br />

differentiation (gametangio-game<strong>to</strong>gamy) has


INTRODUCTION<br />

133<br />

Fig 6.5 Sexual reproduction in Zygorhizidium plank<strong>to</strong>nicum. (a) Empty donor gametangium <strong>to</strong> the left connected by a<br />

conjugation tube <strong>to</strong> a mature resting spore. (b) Near-median section of a fully formed meiosporangium which has developed<br />

from a germinating resting spore.The donor gametangium is on the right. Scale bar ¼ 4 mm. After Doggett and Porter (1996).<br />

been reported in species of Rhizophydium (Karling,<br />

1977); this involves copulation between the<br />

gametangium of a rhizoid-forming thallus and<br />

a motile gamete that encysts directly on the<br />

gametangium.<br />

Generally the product of sexual reproduction<br />

is a resting spore or resting sporangium with<br />

thick walls, but it is known that thick-walled<br />

sporangia may also develop asexually and in<br />

many chytrids sexual reproduction has not been<br />

described and possibly does not occur. Resting<br />

sporangia of some chytrids may remain viable<br />

for many years.<br />

6.1.4 Classification and evolution<br />

Fossil chytrids have been reported from the<br />

400 million-year-old Rhynie chert, a site known<br />

for the discovery of fossil remains of the<br />

earliest known vascular land plants. Clusters<br />

of holocarpic, endobiotic thalli resembling the<br />

present day Olpidium have been found inside<br />

cells of a coenobial alga preserved within the<br />

hollow axes of a vascular plant, and epibiotic<br />

sporangia with endobiotic rhizoids have been<br />

seen attached <strong>to</strong> meiospores of a vascular plant,<br />

much like those of extant chytrids like<br />

Rhizophydium which grow on pollen grains<br />

(Taylor et al., 1992). Chytrid-like fossils have also<br />

been found in strata of the 340 million-year-old<br />

Pennsylvanian (Carboniferous) era (Millay &<br />

Taylor, 1978) and from the more recent Eocene<br />

strata (Bradley, 1967).<br />

Formerly thought <strong>to</strong> have an affinity for<br />

the Oomycota, Hyphochytriomycota or protists,<br />

the Chytridiomycota are now accepted as<br />

members of the true fungi, the Eumycota. They<br />

are probably ancestral <strong>to</strong> other groups of true<br />

fungi, especially the Zygomycota (Cavalier-Smith,<br />

1987, 2001; D. J. S. Barr, 2001). The inclusion of the<br />

chytrids in the Eumycota is supported by several<br />

DNA-based phylogenetic analyses (e.g. Bowman<br />

et al., 1992; James et al., 2000), but the delimitation<br />

of orders within the Chytridiomycota is still<br />

problematic. Particularly puzzling is the grouping<br />

of the Blas<strong>to</strong>cladiales with the Zygomycota on<br />

the basis of 18S ribosomal DNA sequences<br />

(see Fig. 1.26).<br />

D. J. S. Barr (2001) and Kirk et al. (2001) have<br />

classified the Chytridiomycota in<strong>to</strong> five orders<br />

(Table 6.1) but the details of their distinguishing<br />

features need not concern us here. We shall<br />

study examples from each order.


134 CHY TRIDIOMYCOTA<br />

Table 6.1. Orders of Chytridiomycota following D. J. S. Barr (2001) and Kirk et al. (2001).<br />

Order<br />

Number of<br />

described taxa<br />

Examples<br />

Chytridiales<br />

(see Section 6.2)<br />

Spizellomycetales<br />

(see Section 6.3)<br />

Neocallimastigales<br />

(see Section 6.4)<br />

Blas<strong>to</strong>cladiales<br />

(see Section 6.5)<br />

Monoblepharidales<br />

(see Section 6.6)<br />

80 genera Cladochytrium, Nowakowskiella, Rhizophydium,<br />

600 spp.<br />

Synchytrium<br />

13 genera Olpidium, Rhizophlyctis<br />

86 spp.<br />

5genera<br />

16 spp.<br />

Anaeromyces,Caecomyces, Neocallimastix,<br />

Orpinomyces, Piromyces<br />

14 genera Allomyces, Blas<strong>to</strong>cladiella,Coelomomyces,<br />

179 spp.<br />

Physoderma<br />

4genera<br />

19 spp.<br />

Gonapodya, Monoblepharella, Monoblepharis<br />

6.2 Chytridiales<br />

This is by far the largest order, comprising more<br />

than 50% of the <strong>to</strong>tal number of chytrids<br />

described <strong>to</strong> date. It is difficult <strong>to</strong> characterize<br />

members of the Chytridiales because they lack<br />

any specific features by which species have been<br />

assigned <strong>to</strong> the other four orders. The classification<br />

of the Chytridiales has traditionally been<br />

based on thallus morphology (Sparrow, 1973)<br />

but, as pointed out by D. J. S. Barr (2001), this is<br />

unsatisfac<strong>to</strong>ry because of the great variability in<br />

thallus organization shown by the same fungus<br />

growing on its natural substratum and in<br />

culture. Future systems of classification will be<br />

based on zoospore ultrastructure and the<br />

comparison of several different types of DNA<br />

sequences, but <strong>to</strong>o few examples have yet been<br />

studied <strong>to</strong> provide a definitive framework.<br />

Because of this we shall study genera which<br />

illustrate the range of morphology, life cycles<br />

and ecology of the Chytridiales without attempting<br />

<strong>to</strong> place them in<strong>to</strong> families.<br />

6.2.1 Synchytrium<br />

In this genus the thallus is endobiotic and<br />

holocarpic, and at reproduction it may become<br />

converted directly in<strong>to</strong> a group (sorus) of<br />

sporangia, or <strong>to</strong> a prosorus which later gives<br />

rise <strong>to</strong> a sorus of sporangia. Alternatively the<br />

thallus may turn in<strong>to</strong> a resting spore which<br />

can function either directly as a sporangium<br />

and give rise <strong>to</strong> zoospores, or as a prosorus. The<br />

zoospores are of the characteristic chytrid type<br />

(Lange & Olson, 1978). Sexual reproduction is<br />

by copulation of isogametes, resulting in the<br />

formation of thalli which develop in<strong>to</strong> thickwalled<br />

resting spores. Synchytrium includes<br />

about 120 species which are biotrophic parasites<br />

of flowering plants. Some species parasitize only<br />

a narrow range of hosts, e.g. S. endobioticum on<br />

Solanaceae, but others, e.g. S. macrosporum, have<br />

a wide host range (Karling, 1964). Most species<br />

are not very destructive <strong>to</strong> the host plant but<br />

stimulate the formation of galls on leaves,<br />

stems and fruits.<br />

Synchytrium endobioticum<br />

This is the cause of wart disease affecting<br />

cultivated pota<strong>to</strong>es and some wild species of<br />

Solanum. It is a biotrophic pathogen which has<br />

not yet been successfully cultured outside living<br />

host cells. Wart disease is now distributed<br />

throughout the main pota<strong>to</strong>-growing regions<br />

of the world, especially in mountainous areas<br />

and those with a cool, moist climate. Lange<br />

(1987) has given practical details of techniques<br />

for studying the fungus but in most European<br />

countries handling of living material by


CHYTRIDIALES<br />

135<br />

unlicensed workers is illegal. Diseased tubers<br />

bear dark brown cauliflower-like excrescences.<br />

Galls may also be formed on the aerial shoots,<br />

and they are then green with convoluted leaf-like<br />

masses of tissue (the leafy gall stage; Plates 3a,b).<br />

Heavily infected tubers may have a considerable<br />

proportion of their tissues converted <strong>to</strong> warts.<br />

The yield of saleable pota<strong>to</strong>es from a heavily<br />

infected crop may be less than the actual weight<br />

of the seed pota<strong>to</strong>es planted. The disease is thus<br />

potentially a serious one, but fortunately varieties<br />

of pota<strong>to</strong>es are available which are immune<br />

from the disease, so that control is practicable.<br />

The possible life cycle of S. endobioticum is<br />

summarized in Fig. 6.6.<br />

The dark warts on the tubers are galls in<br />

which the host cells have been stimulated <strong>to</strong><br />

divide by the presence of the fungus. Many<br />

of the host cells contain resting spores which<br />

are more or less spherical cells with thick dark<br />

brown walls folded in<strong>to</strong> plate-like extensions<br />

(see Fig. 6.7a). The resting spores are released<br />

by the decay of the warts and may remain alive<br />

in the soil for over 40 years (Laidlaw, 1985).<br />

The outer wall (exospore) bursts open by an<br />

irregular aperture and the endospore balloons<br />

out <strong>to</strong> form a vesicle within which a single<br />

sporangium differentiates (Kole, 1965; Sharma<br />

& Cammack, 1976; Hampson et al., 1994). Thus<br />

the resting spore functions as a prosporangium<br />

on germination. Germination of the resting<br />

spore may occur spontaneously but can be stimulated<br />

by passage through snails. It is presumed<br />

that abrasion and digestion of the spore wall<br />

Fig 6.6 Schematic outline of the probable life cycle of Synchytrium endobioticum. Haploid and diploid nuclei are represented by<br />

small empty and larger split circles, respectively. Key events in the life cycle are plasmogamy (P), karyogamy (K) and meiosis (M).<br />

Resting spores within a warted pota<strong>to</strong> contain a single nucleus which undergoes meiosis upon germination. Haploid zoospores<br />

are released from a single sporangium. If two zoospores pair up, a zygote is formed and penetration of a pota<strong>to</strong> cell gives rise <strong>to</strong><br />

a diploid thallus and, ultimately, a resting spore. Diploid infections cause host hyperplasia visible as the pota<strong>to</strong> wart symp<strong>to</strong>ms.<br />

If a zoospore infects in the haploid state, a haploidprosorus (summer spore) is formed, and hypertrophy of the infected and adjacent<br />

host cells ensues. A sorus of several sporangia is ultimately produced, with each sporangium releasing a fresh crop of haploid<br />

zoospores. Synchytrium endobioticum appears <strong>to</strong> be homothallic.


136 CHY TRIDIOMYCOTA<br />

Fig 6.7 Synchytrium endobioticum.<br />

(a) Resting spores in section of wart.<br />

(b) Germinating resting spore showing<br />

the formation of a vesicle containing<br />

a single globose sporangium (after Kole,<br />

1965). (c) Section of infected host cell<br />

containing a prosorus.The prosorus<br />

is extruding a vesicle. Note the hypertrophy<br />

of the infected cell and adjacent<br />

uninfected cells. (d) Cleavage of vesicle<br />

contents <strong>to</strong> form zoosporangia. (e) Two<br />

extruded zoosporangia. (f) Zoospores.<br />

(g) Rosette of hypertrophied pota<strong>to</strong> cells<br />

as seen from the surface.The outline<br />

of the infected host cell is shown dotted.<br />

(h) Young resting sporangium resulting<br />

from infection by a zygote. Note that<br />

the infected cell lies beneath the<br />

epidermis due <strong>to</strong> division of the<br />

host cells.<br />

in the snail gut causes breakdown of the thick<br />

wall which contains chitin and branched-chain<br />

wax esters, so overcoming dormancy related<br />

<strong>to</strong> the impermeability of the wall (Hampson<br />

et al., 1994).<br />

The zoospores are capable of swimming for<br />

about two hours in the soil water. If they alight<br />

on the surface of a pota<strong>to</strong> ‘eye’ or some other<br />

part of the pota<strong>to</strong> shoot such as a s<strong>to</strong>lon or<br />

a young tuber before its epidermis is suberized,<br />

they come <strong>to</strong> rest and withdraw their flagellum.<br />

During penetration, the contents of the zoospore<br />

cyst are transferred <strong>to</strong> the host cell whilst the<br />

cyst wall remains attached <strong>to</strong> the outside. When<br />

a dormant ‘eye’ is infected, dormancy may<br />

be broken and the tuber may begin <strong>to</strong> sprout. If<br />

the pota<strong>to</strong> variety is susceptible <strong>to</strong> the disease,<br />

the small fungal thallus inside the host cell<br />

will enlarge. The infected host cell as well as<br />

surrounding cells also enlarge so that a rosette<br />

of hypertrophied cells surrounds a central<br />

infected cell (Fig. 6.7c). The walls of these cells<br />

adjacent <strong>to</strong> the infected cell are often thickened<br />

and assume a dark brown colour. The infected<br />

cell remains alive for some time but eventually<br />

it dies. The pathogen thallus passes <strong>to</strong> the<br />

bot<strong>to</strong>m of the host cell, enlarges and becomes<br />

spherical. A double-layered chitinous wall which


CHYTRIDIALES<br />

137<br />

is golden brown in colour is secreted around<br />

the thallus, now termed a prosorus or summer<br />

spore. Further development of the prosorus<br />

involves the protrusion of the inner wall through<br />

a pore in the outer wall, and its expansion as<br />

a vesicle which enlarges upwards and fills the<br />

upper half of the host cell (Fig. 6.7c). The<br />

cy<strong>to</strong>plasmic contents of the prosorus including<br />

the single nucleus are transferred <strong>to</strong> the vesicle.<br />

The process is quite rapid and can be completed<br />

in about 4 h. During its passage in<strong>to</strong> the vesicle<br />

the nucleus may divide, and mi<strong>to</strong>ses continue<br />

so that the vesicle contains about 32 nuclei.<br />

At this stage the cy<strong>to</strong>plasmic contents of the<br />

vesicle become cleaved in<strong>to</strong> about 4 9 sporangia<br />

(Fig. 6.7d), forming a sorus. After the deposition<br />

of sporangial walls, further nuclear divisions<br />

occur in each sporangium, and finally each<br />

nucleus with its surrounding mass of cy<strong>to</strong>plasm<br />

becomes differentiated <strong>to</strong> form a zoospore. As<br />

the sporangia ripen, they absorb water and<br />

swell, causing the host cell containing them <strong>to</strong><br />

burst open. Meanwhile, division of the host<br />

cells underlying the rosette has been taking<br />

place, and enlargement of these cells pushes<br />

the sporangia out on<strong>to</strong> the surface of the host<br />

tissue (Fig. 6.7e). The sporangia swell if water<br />

is available and burst open by means of a small<br />

slit through which the zoospores escape. There<br />

may be as many as 500 600 zoospores in a single<br />

large sporangium. The zoospores resemble those<br />

derived from resting sporangia and are capable<br />

of initiating further asexual cycles of reproduction<br />

throughout spring and early summer.<br />

Sometimes several zoospores succeed in penetrating<br />

a single cell so that it contains several<br />

fungal pro<strong>to</strong>plasts.<br />

Alternatively, zoospores may function as<br />

gametes, fusing in pairs (or occasionally in<br />

groups of three or four) <strong>to</strong> form zygotes which<br />

retain their flagella and swim actively for a time.<br />

Since zoospores acting as gametes do not differ<br />

in size and shape, copulation can be described<br />

as isogamous. However, the gametes may differ<br />

physiologically. Curtis (1921) has suggested that<br />

fusion may not occur between zoospores derived<br />

from a single sporangium, but only between<br />

zoospores from separate sporangia. Köhler (1956)<br />

has claimed that the zoospores are at first<br />

sexually neutral. Later they mature and become<br />

capable of copulation. Maturation may occur<br />

either outside the sporangia or within, so that<br />

in over-ripe sporangia the zoospores are capable<br />

of copulation on release. At first the zoospores<br />

are ‘male’, and swim actively. Later the swarmers<br />

become quiescent (‘female’) and probably secrete<br />

a substance which attracts ‘male’ gametes.<br />

After swimming by means of its two flagella,<br />

the zygote encysts on the surface of the host<br />

epidermis and penetration may then follow by<br />

a process essentially similar <strong>to</strong> zoospore penetration.<br />

Multiple infections by several zygotes<br />

penetrating a single host cell can also occur.<br />

Nuclear fusion occurs in the young zygote before<br />

penetration.<br />

The results of zygote infections differ from<br />

infection by zoospores. The host cell reacts <strong>to</strong><br />

zoospore infection by undergoing hypertrophy,<br />

i.e. increase in cell volume, and adjacent cells<br />

also enlarge <strong>to</strong> form the characteristic rosette<br />

which surrounds the resulting prosorus. In<br />

contrast, when a zygote infects, the host cell<br />

undergoes hyperplasia, i.e. repeated cell division.<br />

The pathogen lies <strong>to</strong>wards the bot<strong>to</strong>m of the<br />

host cell, adjacent <strong>to</strong> the host nucleus, and cell<br />

division occurs in such a way that the fungal<br />

pro<strong>to</strong>plast is located in the innermost daughter<br />

cell. As a result of repeated divisions of the<br />

host cells, the typical gall-like pota<strong>to</strong> warts are<br />

formed and fungal pro<strong>to</strong>plasts may be buried<br />

several cell layers deep beneath the epidermis<br />

(see Fig. 6.7h). During these divisions of the<br />

host tissue the zygote enlarges and becomes<br />

surrounded by a two-layered wall, a thick outer<br />

layer which eventually becomes dark brown in<br />

colour and is thrown in<strong>to</strong> folds or ridges which<br />

appear as spines in section, and a thin hyaline<br />

inner wall surrounding the granular cy<strong>to</strong>plasm<br />

(Lange & Olson, 1981). The host cell eventually<br />

dies and some of its contents are deposited on<br />

the outer wall of the resting sporangium,<br />

forming the characteristic brown ridges. During<br />

its development the resting spore remains uninucleate.<br />

Resting spores are released in<strong>to</strong> the<br />

soil and are capable of germination within<br />

about 2 months. Before germination, the nucleus<br />

divides repeatedly <strong>to</strong> form the nuclei of the<br />

zoospores whose further development has


138 CHY TRIDIOMYCOTA<br />

already been described. It has been claimed that<br />

the zygote and the young resting spore<br />

are diploid, and it has been assumed that meiosis<br />

occurs during germination of the resting sporangia<br />

prior <strong>to</strong> the formation of zoospores, so<br />

that these zoospores, the prosori and the soral<br />

zoospores are also believed <strong>to</strong> be haploid.<br />

These assumptions seem plausible in the light<br />

of knowledge of the life his<strong>to</strong>ry and cy<strong>to</strong>logy<br />

of other species (e.g. Lingappa, 1958b), and an<br />

essentially similar life cycle has been described<br />

for S. lagenariae and S. trichosanthidis, parasitic on<br />

Cucurbitaceae, which differ from S. endobioticum<br />

in that their resting spores function as prosori<br />

instead of prosporangia (Raghavendra Rao &<br />

Pavgi, 1993).<br />

Control of wart disease<br />

Control is based largely on the breeding of<br />

resistant varieties of pota<strong>to</strong>. It was discovered<br />

that certain varieties such as Snowdrop were<br />

immune from the disease and could be planted<br />

on land heavily infected with Synchytrium without<br />

developing warts. Following this discovery,<br />

plant breeders have developed a number of<br />

immune varieties such as Maris Piper. However,<br />

some pota<strong>to</strong> varieties that are susceptible <strong>to</strong><br />

the disease are still widely grown, including the<br />

popular King Edward. In most countries where<br />

wart disease occurs, legislation has been introduced<br />

requiring that only approved immune<br />

varieties be planted on land where wart disease<br />

has been known <strong>to</strong> occur, and prohibiting the<br />

movement and sale of diseased material. Within<br />

the British Isles, the growing of immune varieties<br />

on infested land has prevented the spread of<br />

the disease, and it is now confined <strong>to</strong> a small<br />

number of foci in the West Midlands, northwest<br />

England and mid and south Scotland. It has<br />

also persisted in Newfoundland. The majority of<br />

the outbreaks are found in allotments, gardens<br />

and smallholdings.<br />

The reaction of immune varieties <strong>to</strong> infection<br />

varies (Noble & Glynne, 1970). In some cases<br />

when ‘immune’ varieties are exposed <strong>to</strong> a heavy<br />

inoculum load of S. endobioticum in the labora<strong>to</strong>ry,<br />

they may become slightly infected, but<br />

infection is often confined <strong>to</strong> the superficial<br />

tissues which are soon sloughed off. In the<br />

field such slight infections would probably pass<br />

unnoticed. Occasionally infections of certain<br />

pota<strong>to</strong> varieties may result in the formation<br />

of resting spores, but without the formation<br />

of noticeable galls. Penetration of the parasite<br />

seems <strong>to</strong> occur in all pota<strong>to</strong> varieties, but when<br />

a cell of an immune variety is penetrated it<br />

may die within a few hours, and since the fungus<br />

is a biotrophic parasite, further development is<br />

checked. In other cases the parasite may persist<br />

in the host cell for up <strong>to</strong> 2 3 days, apparently<br />

showing normal development, but after this<br />

time the fungal thallus undergoes disorganization<br />

and disappears from the host cell.<br />

Unfortunately, it has been discovered that<br />

new physiological races (or pathotypes) of<br />

the pathogen have arisen, capable of attacking<br />

pota<strong>to</strong> varieties previously thought <strong>to</strong> be<br />

immune. About 20 pathotypes are now known,<br />

and the implications are obvious. Unless their<br />

spread can be prevented, much of the work of<br />

pota<strong>to</strong> plant breeders over the past century will<br />

have <strong>to</strong> be started all over again.<br />

Other methods of control are less satisfac<strong>to</strong>ry.<br />

Attempts <strong>to</strong> kill the resting spores of the<br />

fungus in the soil have been made, but this is<br />

a costly and difficult process, requiring largescale<br />

fungicide applications <strong>to</strong> the soil. Copper<br />

sulphate or ammonium thiocyanate have been<br />

applied in the past at amounts of up <strong>to</strong> 1 <strong>to</strong>n<br />

acre 1 , and local treatment with mercuric<br />

chloride or with formaldehyde and steam has<br />

been used <strong>to</strong> eradicate foci of infection<br />

(Hampson, 1988). Control measures based on<br />

the use of resistant varieties seem more satisfac<strong>to</strong>ry.<br />

An interesting method of control developed<br />

in Newfoundland is the use of crabshell meal<br />

placed above seed tubers at the time of planting.<br />

This technique has resulted in significant<br />

and sometimes complete control (Hampson &<br />

Coombes, 1991) which may be due <strong>to</strong> selective<br />

enhancement of chitinolytic soil micro-organisms<br />

degrading the chitinous walls of the resting<br />

spores of S. endobioticum.<br />

Other species of Synchytrium<br />

Not all species of Synchytrium show the same kind<br />

of life cycle as S. endobioticum. Synchytrium fulgens,<br />

a parasite of Oenothera, resembles S. endobioticum


CHYTRIDIALES<br />

139<br />

in that both summer spores and resting spores<br />

are formed (Lingappa, 1958a,b), but in this<br />

species the zoospores from resting sporangia<br />

can also function as gametes and give rise<br />

directly <strong>to</strong> zygote infections from which further<br />

resting spores arise (Lingappa, 1958b). It has<br />

been suggested that the same phenomenon may<br />

occasionally occur in S. endobioticum. InS. taraxaci<br />

parasitic on Taraxacum (Fig. 6.8; Plate 3c), as<br />

well as a number of other Synchytrium spp., the<br />

mature thallus does not function as a prosorus<br />

but cleaves directly <strong>to</strong> form a sorus of sporangia,<br />

and the resting spore also gives rise <strong>to</strong> zoospores<br />

directly. In some species, e.g. S. aecidioides,<br />

resting sporangia are unknown, whilst in others,<br />

e.g. S. mercurialis, a common parasite on leaves<br />

and stems of Mercurialis perennis (Fig. 6.9), only<br />

resting sporangia are known and summer sporangial<br />

sori do not occur. Mercurialis plants<br />

collected from March <strong>to</strong> June often show<br />

yellowish blisters on leaves and stems. The<br />

blisters are galls made up of one or two layers<br />

of hypertrophied cells mostly lacking chlorophyll,<br />

surrounding the Synchytrium thallus<br />

during its maturation <strong>to</strong> form a resting<br />

sporangium. In this species the resting sporangium<br />

functions as a prosorus during the following<br />

spring. The undivided contents are extruded<br />

in<strong>to</strong> a spherical sac which becomes cleaved in<strong>to</strong> a<br />

sorus containing as many as 120 sporangia from<br />

which zoospores arise. The variations in the life<br />

his<strong>to</strong>ries of the various species of Synchytrium<br />

form a useful basis for classifying the genus<br />

(Karling, 1964).<br />

6.2.2 Rhizophydium<br />

Rhizophydium is a large, cosmopolitan genus of<br />

about 100 species (Sparrow, 1960) which grow<br />

in soil, freshwater and the sea. The thallus is<br />

eucarpic, with a globose epibiotic zoosporangium<br />

which develops from the zoospore cyst,<br />

and endobiotic rhizoids which penetrate the<br />

host. Whilst some species are saprotrophic,<br />

others are biotrophic pathogens of algae and<br />

can cause severe epidemics of freshwater phy<strong>to</strong>plank<strong>to</strong>n.<br />

Saprotrophic forms such as R. pollinispini<br />

and R. sphaerocarpon colonize pollen grains<br />

and are easily isolated by sprinkling pollen on<strong>to</strong><br />

the surface of water overlying soil (Fig. 6.10).<br />

Within 3 days, sporangia with exit papillae are<br />

Fig 6.8 Synchytrium taraxaci.<br />

(a) Undivided thallus in epidermal cell of<br />

scape of Taraxacum.Outline of host cell<br />

shown dotted. (b) Section of Taraxacum<br />

scape showing thallus divided in<strong>to</strong> a sorus<br />

of sporangia. (c) A sorus of<br />

sporangia seen from above.<br />

Two sporangia are releasing zoospores.<br />

(d) A ripe sporangium. (e) Sporangium<br />

releasing zoospores. (f) Zoospores and<br />

zygotes.The triflagellate zoospore<br />

probably arose by incomplete<br />

separation of zoospore initials.<br />

(g) Section of host leaf showing a resting<br />

sporangium. (a e) and (g) <strong>to</strong> same scale.


140 CHYTRIDIOMYCOTA<br />

Fig 6.9 Synchytrium mercurialis.<br />

(a) Section of stem of Mercurialis<br />

perennis showing hypertrophied cells<br />

surrounding a resting sporangium.<br />

(b) Germination of a resting<br />

sporangium <strong>to</strong> release a sorus of<br />

zoosporangia.Thus in S. mercurialis<br />

the resting sporangium functions as<br />

a prosorus (after Fischer,1892).<br />

Fig 6.10 Pine pollen grains<br />

colonized by Rhizophydium sp.<br />

(a) The rhizoid system attaching<br />

the epibiotic sporangium <strong>to</strong> the<br />

colonized pollen grain. (b) Mature<br />

sporangium; the cy<strong>to</strong>plasm has<br />

become cleaved in<strong>to</strong> numerous<br />

zoospores.<br />

found in crude cultures on pine pollen. The<br />

zoospores are at first released in<strong>to</strong> a hyaline<br />

vesicle which soon dissolves, allowing them <strong>to</strong><br />

swim away. Gauriloff and Fuller (1987) have<br />

outlined techniques for growing R. sphaerocarpon<br />

in pure culture. This species can also grow<br />

parasitically on filaments of the green alga<br />

Spirogyra.<br />

Douglas Lake (Michigan, USA) is surrounded<br />

by conifers shedding pollen which floats on<br />

the lake and becomes colonized by Rhizophydium<br />

spp. Using the MPN (most probable number)<br />

technique, Ulken and Sparrow (1968) have<br />

estimated that the number of chytrid propagules<br />

in the surface waters (epilimnion) can rise <strong>to</strong><br />

over 900 l 1 by late June. Some infected pollen<br />

grains sink through the hypolimnion <strong>to</strong> the mud<br />

at the floor of the lake. It is thought likely that<br />

these develop resting sporangia which survive<br />

the winter and provide inoculum <strong>to</strong> start off<br />

colonization of new pollen deposits in the<br />

following season.<br />

Rhizophydium plank<strong>to</strong>nicum<br />

This species is the best-studied chytrid phy<strong>to</strong>plank<strong>to</strong>n<br />

parasite. It is a biotrophic pathogen of


CHYTRIDIALES<br />

141<br />

the dia<strong>to</strong>m Asterionella formosa, an inhabitant of<br />

eutrophic lakes. This alga forms cartwheel-like<br />

colonies, the dia<strong>to</strong>m frustules making up the<br />

spokes, cemented <strong>to</strong>gether by mucilage pads at<br />

the hub of the wheel. Rhizophydium plank<strong>to</strong>nicum<br />

may form one <strong>to</strong> many thalli on each host<br />

cell (Fig. 6.11a). Dual cultures of the host and<br />

parasite have been established (Canter &<br />

Jaworski, 1978) and from such cultures a detailed<br />

picture of infection, development and zoospore<br />

structure has been built up (Beakes et al., 1993).<br />

Zoospores are attracted <strong>to</strong> the alga and encyst<br />

on it, forming monocentric rhizoidal thalli.<br />

The rhizoids penetrate between the girdle lamellae<br />

of the host (Fig. 6.11b). The rhizoids may<br />

extend throughout the whole length of the<br />

host cell and infection is often accompanied<br />

by loss of pho<strong>to</strong>synthetic pigment, failure of<br />

cells <strong>to</strong> divide, and ultimately early death of<br />

the host cell. The zoospore is uninucleate and the<br />

nucleus is retained within the zoospore cyst,<br />

the rhizoids being devoid of nuclei. The zoospore<br />

cyst enlarges <strong>to</strong> form the sporangium.<br />

Synchronous nuclear divisions result in the<br />

formation of several nuclei lying within the<br />

cy<strong>to</strong>plasm, followed by the development of<br />

cleavage furrows which divide up the sporangial<br />

contents in<strong>to</strong> zoospores. A septum develops at the<br />

base of the sporangium and, prior <strong>to</strong> cleavage, the<br />

upper part of the sporangium wall develops a<br />

thickened apical papilla which balloons out <strong>to</strong><br />

form a vesicle in<strong>to</strong> which the immobile zoospores<br />

are released. The complete cycle from infection <strong>to</strong><br />

zoospore release depends on temperature<br />

and can be as short as 2 3 days. About 1 30<br />

zoospores may be formed in a sporangium<br />

depending on the state of the host cells, in turn<br />

affected by external physical and chemical conditions.<br />

Breakdown of the vesicle allows the<br />

zoospores <strong>to</strong> swim away. No resting stage has<br />

been described for R. plank<strong>to</strong>nicum.<br />

A striking feature of the zoospore ultrastructure<br />

is the presence of several paracrystalline<br />

bodies near the nucleus in the peripheral part<br />

of the cy<strong>to</strong>plasm (Beakes et al., 1993). They consist<br />

of parallel arrays of regularly arranged crystals<br />

interconnected <strong>to</strong> each other with fibrous<br />

material. They appear late in sporangial development<br />

but disappear following encystment<br />

of zoospores. Similar structures have been<br />

reported from the zoospores of a few other<br />

Chytridiomycota, but their composition and<br />

function are unknown.<br />

There have been several studies on the<br />

ecology of Asterionella subjected <strong>to</strong> parasitism by<br />

R. plank<strong>to</strong>nicum (see Canter & Lund, 1948, 1953;<br />

Fig 6.11 Rhizophydium plank<strong>to</strong>nicum<br />

growing parasitically on the<br />

frustules of the colonial dia<strong>to</strong>m<br />

Asterionella formosa. (a) Heavily<br />

infected colony from a dual-clone<br />

culture showing encysted<br />

zoospores. (b) Scanning electron<br />

micrograph of Asterionella cells<br />

showing heavy infection and<br />

zoospore cysts which have<br />

germinated and penetrated the<br />

host cells via the girdle lamellae.<br />

From Beakes et al. (1993), with<br />

permission from Elsevier; original<br />

images kindly provided by G.W.<br />

Beakes.


142 CHYTRIDIOMYCOTA<br />

Canter & Jaworski, 1981; van Donk & Bruning,<br />

1992). Asterionella is also parasitized by two<br />

other chytrids, Zygorhizidium plank<strong>to</strong>nicum and<br />

Z. affluens, and some of the early studies in freshwater<br />

lakes may well have included a mixture<br />

of species.<br />

Studies on the epidemiology of infection of<br />

Asterionella by R. plank<strong>to</strong>nicum in lakes have<br />

shown that there are peak periods of Asterionella<br />

population density both in spring and in autumn,<br />

related <strong>to</strong> the availability of dissolved nutrients,<br />

water temperature, thermal stratification and its<br />

breakdown, daylength and light intensity.<br />

Asterionella cells infected with Rhizophydium can<br />

occur throughout the year, but epidemics in<br />

which a high proportion of cells are infected only<br />

occur at concentrations of around 10 host cells<br />

ml 1 (Holfeld, 1998). Interpretation of the conditions<br />

conducive <strong>to</strong> the occurrence of epidemics<br />

has been aided by experiments using dual<br />

cultures of pathogen and host in which effects<br />

such as light intensity, temperature and phosphorus<br />

concentration have been varied (van Donk<br />

& Bruning, 1992). The effects of light are complex.<br />

Although Rhizophydium zoospores are not pho<strong>to</strong>tropic,<br />

they are quiescent and incapable of<br />

infection in the dark or at low light intensity.<br />

Experiments by Canter and Jaworski (1981) have<br />

indicated that a light intensity below 200 lx is<br />

inadequate for zoospore settlement on host cells.<br />

In light-limited cultures of Asterionella, the sporangia<br />

of the pathogen and hence the number of<br />

zoospores produced are smaller than when light<br />

is not limiting (Bruning, 1991a). Similarly, zoospore<br />

production is also reduced when the<br />

concentration of phosphorus limits growth of<br />

the host (Bruning, 1991b). Temperature affects<br />

the rate of sporangium development and the<br />

size of sporangia, with maximum dimensions<br />

at 2°C at fairly high light intensities (Bruning,<br />

1991a). It also affects the duration of swimming of<br />

zoospores and therefore their infective lifetime<br />

which can vary from about 10 days at 3°C <strong>to</strong> only<br />

2 days at 20°C. Epidemic development may result<br />

from a combination of fac<strong>to</strong>rs and there is a<br />

remarkable interaction between the effects of<br />

light intensity and temperature (Bruning, 1991c).<br />

At higher temperatures, optimal conditions for<br />

epidemic development occur at high light<br />

intensities, but at temperatures below 5 6°C<br />

epidemic development is encouraged by lower<br />

light intensities. This may explain why, in nature,<br />

epidemics can occur both in summer (high light<br />

intensity, high temperature) and winter (low<br />

light intensity, low temperature).<br />

Rhizophydium plank<strong>to</strong>nicum is a specialized<br />

parasite infecting only Asterionella. It is more<br />

compatible with certain clones of host cells<br />

than others, and cells from incompatible clones<br />

show hypersensitivity, undergoing rapid death<br />

following infection (Canter & Jaworski, 1979).<br />

6.2.3 Cladochytrium<br />

There are about a dozen species of Cladochytrium<br />

(Sparrow, 1960) which are widespread saprotrophs,<br />

mostly of aquatic plant debris. The<br />

thallus is eucarpic and polycentric and the<br />

vegetative system may bear intercalary swellings<br />

and septate turbinate cells (sometimes termed<br />

spindle organs). The sporangia are inoperculate.<br />

Cladochytrium replicatum is a common representative<br />

in decaying pieces of aquatic vegetation<br />

and can be distinguished from other chytrids<br />

by the bright orange lipid droplets found in the<br />

sporangia. It is frequently isolated if moribund<br />

aquatic vegetation is placed in a dish of water<br />

and baited with boiled grass leaves or cellulosic<br />

materials such as dialysis tubing. Lucarotti (1987)<br />

has given details of its isolation and growth<br />

in culture. The bright orange sporangia which<br />

are visible under a dissecting microscope<br />

appear on baits within about 5 days, arising<br />

from an extensively branched hyaline rhizomycelium<br />

bearing two-celled intercalary swellings.<br />

Sporangium development is encouraged by<br />

exposure <strong>to</strong> light. On release from the sporangium,<br />

the zoospores each contain a single orange<br />

lipid droplet and bear a single posterior flagellum.<br />

Lucarotti (1981) has described the fine<br />

structure of the zoospore. After swimming for<br />

a short time, the zoospore attaches itself <strong>to</strong><br />

the surface of the substratum and puts out<br />

usually a single germ tube which can penetrate<br />

the tissues of the host plant. The germ tube<br />

expands <strong>to</strong> form an elliptical or cylindrical<br />

turbinate cell which is often later divided in<strong>to</strong><br />

two by a transverse septum (Fig. 6.12d). The


CHYTRIDIALES<br />

143<br />

zoospore is uninucleate and during germination<br />

the single nucleus is transferred <strong>to</strong> the<br />

swollen turbinate cell which becomes a vegetative<br />

centre from which rhizoids are put out<br />

which in turn produce further turbinate cells<br />

(see Figs. 6.12b,d). Nuclear division is apparently<br />

confined <strong>to</strong> the turbinate cells, and although<br />

nuclei are transported through the rhizoidal<br />

system they are not resident there. The thallus<br />

so established branches profusely, and at certain<br />

points spherical zoosporangia form,<br />

either terminally or in intercalary positions.<br />

Sometimes one of the cells of a pair of turbinate<br />

cells swells and becomes transformed in<strong>to</strong> a<br />

sporangium. In culture, both cells may<br />

be modified in this way. The spherical <strong>to</strong> pearshaped<br />

zoosporangium undergoes progressive<br />

nuclear division, and the contents of the sporangium<br />

acquire a bright orange colour due<br />

<strong>to</strong> accumulation of lipid droplets containing<br />

the carotenoid lycopene. These lipid reserves are<br />

later found in the zoospores. Cleavage of the<br />

cy<strong>to</strong>plasm <strong>to</strong> form uninucleate zoospore initials<br />

follows. The zoospores escape through a narrow<br />

Fig 6.12 Cladochytriumreplicatum.<br />

(a) Rhizomycelium within the epidermis<br />

of an aquatic plant bearing the<br />

two-celled hyaline turbinate cells and<br />

globose orange zoosporangia. (b)<br />

Rhizomycelium and turbinate cells from<br />

a culture. (c) Zoosporangia from a<br />

two-week-old culture. One<br />

zoosporangium has released zoospores,<br />

each of which contains a bright orangecoloured<br />

globule. (d) Germinating<br />

zoospores on boiled wheat leaves.<br />

The empty zoospore cysts are spherical.<br />

The germ tubes have expanded <strong>to</strong> form<br />

turbinate cells. (e) A zoosporangium<br />

which has proliferated internally <strong>to</strong> form<br />

a second sporangium. (f) Rhizomycelium<br />

within a boiled wheat leaf bearing a<br />

thick-walled, spiny resting sporangium.


144 CHYTRIDIOMYCOTA<br />

exit tube which penetrates <strong>to</strong> the exterior of<br />

the substratum and becomes mucilaginous at<br />

the tip. There is no operculum. Sometimes<br />

zoosporangia may proliferate internally, a new<br />

zoosporangium being formed inside the wall of<br />

an empty one. Resting sporangia with thicker<br />

walls and a more hyaline cy<strong>to</strong>plasm are also<br />

formed either terminally or in an intercalary<br />

position on the rhizomycelium. In some cases<br />

the wall of the resting sporangium is reported<br />

<strong>to</strong> be smooth and in others spiny, and it has<br />

been suggested (Sparrow, 1960) that the two<br />

kinds of resting sporangia may belong <strong>to</strong> different<br />

species. However, studies by Willoughby<br />

(1962) of a number of single-spore isolates have<br />

shown that the presence or absence of spines is<br />

a variable character. The contents of the resting<br />

sporangia divide <strong>to</strong> form zoospores which also<br />

have a conspicuous orange droplet, and escape<br />

by means of an exit tube as in the thin-walled<br />

zoosporangia. Whether the resting sporangia are<br />

formed as a result of a sexual process is not<br />

known. Pure cultures of C. replicatum have been<br />

studied by Willoughby (1962), Goldstein (1960)<br />

and Lucarotti (1981). The fungus is heterotrophic<br />

for thiamine. Biotin, while not absolutely<br />

required, stimulates growth. Nitrate and sulphate<br />

are utilized, as are a number of different<br />

carbohydrates; a limited amount of growth takes<br />

place on cellulose.<br />

6.2.4 Nowakowskiella<br />

Species of Nowakowskiella are widespread saprotrophs<br />

in soil and on decaying aquatic plant<br />

debris, and can be obtained by baiting aquatic<br />

plant remains in water with boiled grass<br />

leaves, cellophane, dialysis tubing and the like.<br />

Nowakowskiella elegans is often encountered in<br />

such material, and pure cultures can be obtained<br />

and grown on cellulosic materials overlying<br />

agar, or directly in liquid culture media<br />

(Emerson, 1958; Johnson, 1977; Lucarotti, 1981;<br />

Lucarotti & Wilson, 1987). In culture, considerable<br />

variation in growth habit and morphology<br />

can result from changing the concentration of<br />

nutrients and the availability of water (Johnson,<br />

1977). In boiled grass leaves the fungus forms an<br />

extensive rhizomycelium with turbinate cells<br />

(Fig. 6.13c). Zoosporangia are formed terminally<br />

or in an intercalary position (Fig. 6.13c) and<br />

are globose or pear-shaped with a subsporangial<br />

swelling (apophysis), and granular or refractile<br />

hyaline contents. At maturity some sporangia<br />

develop a prominent beak, but in others this<br />

is not present. When an operculum becomes<br />

detached, zoospores escape and initially remain<br />

clumped <strong>to</strong>gether at the mouth of the sporangium<br />

(Figs. 6.13b,c). The fine structure of the<br />

zoospore is very similar <strong>to</strong> that of Rhizophydium<br />

but paracrystalline bodies have not been observed<br />

(Lucarotti, 1981). It also has close resemblance<br />

<strong>to</strong> the zoospore ultrastructure of the inoperculate,<br />

polycentric Cladochytrium replicatum.<br />

Yellowish resting sporangia (Fig. 6.13e) have<br />

been described (Emerson, 1958; Johnson, 1977;<br />

Lucarotti & Wilson, 1987). They develop as<br />

spherical <strong>to</strong> fusiform swellings in the rhizomycelium<br />

which become delimited by septa, develop<br />

thick walls and a large central vacuole<br />

surrounded by dense cy<strong>to</strong>plasm with small<br />

spherical lipid droplets. The resting sporangium<br />

is at first binucleate. After nuclear fusion the<br />

diploid nucleus divides meiotically. Further<br />

nuclear divisions are mi<strong>to</strong>tic and the contents<br />

of the resting sporangium cleave in<strong>to</strong> zoospores<br />

which may be released through a papilla in<br />

the sporangium wall. Alternatively, the resting<br />

sporangium may give rise <strong>to</strong> a thin-walled zoosporangium<br />

from which the zoospores are<br />

released, i.e. the resting sporangium may function<br />

as a prosporangium as in some other chytrids<br />

( Johnson, 1977).<br />

In N. profusa, which is probably synonymous<br />

with N. elegans (Johnson, 1977), three kinds of<br />

sporangial dehiscence have been described:<br />

exo-operculate, in which the operculum breaks<br />

away <strong>to</strong> the outside of the sporangium; endooperculate,<br />

in which the operculum remains<br />

within the sporangium; and inoperculate,<br />

where the exit papilla opens without any clearly<br />

defined operculum (Chambers et al., 1967;<br />

Johnson, 1973). Such variations within a single<br />

chytrid strain add emphasis <strong>to</strong> criticisms of<br />

the value of dehiscence as a primary criterion in<br />

classification.<br />

Goldstein (1961) has reported that N. elegans<br />

requires thiamine and can utilize nitrate,


SPIZELLOMYCETALES<br />

145<br />

Fig 6.13 Nowakowskiella elegans.<br />

(a) Polycentric mycelium bearing<br />

zoosporangia. (b) Empty zoosporangia<br />

showing opercula. (c) Mycelium showing<br />

turbinate cells and zoosporangia.<br />

(d) Zoospores from culture. (e) Resting<br />

sporangium from culture.<br />

sulphate and a number of carbohydrates including<br />

cellulose, but cannot utilize starch.<br />

6.3 Spizellomycetales<br />

Members of this order differ from the<br />

Chytridiales in possessing zoospores which<br />

contain more than one lipid droplet and are<br />

capable of limited amoeboid movement. Thalli<br />

are generally monocentric. The order takes its<br />

name from the genus Spizellomyces which in turn<br />

was named in honour of the chytrid pioneer<br />

F. K. Sparrow after Spizella, a genus of North<br />

American sparrows (Barr, 1980). Some 86 species<br />

of Spizellomycetales are currently recognized.<br />

6.3.1 Olpidium<br />

About 30 species of Olpidium are known, but<br />

the genus is in need of revision and possibly


146 CHYTRIDIOMYCOTA<br />

some of the species should be classified elsewhere.<br />

Typical species are holocarpic. Some are<br />

parasitic on fungi and aquatic plants or algae,<br />

or saprotrophic on pollen (Sparrow, 1960).<br />

Others parasitize rotifers (Glockling, 1998),<br />

nema<strong>to</strong>des and their eggs (Tribe, 1977; Barron<br />

& Szijar<strong>to</strong>, 1986), moss pro<strong>to</strong>nemata or leaves<br />

and roots of higher plants (Macfarlane, 1968;<br />

Johnson, 1969). Olpidium bornovanus (¼ O. radicale)<br />

develops on various monocotyledonous and<br />

dicotyledonous plant roots following inoculation<br />

(Lange & Insunza, 1977). Olpidium brassicae is<br />

common on the roots of cabbages, especially<br />

when growing in wet soils, and is also found on<br />

a wide range of unrelated hosts, but some<br />

host specialization has been reported. Both<br />

O. bornovanus and O. brassicae are vec<strong>to</strong>rs of a<br />

number of plant viruses (Barr, 1988; Adams,<br />

1991; Hiruki, 1994; Campbell, 1996) and this<br />

<strong>to</strong>pic is discussed more fully below. Weber and<br />

Webster (2000a) have given practical details of<br />

how <strong>to</strong> grow O. brassicae for observation on<br />

Brassica seedlings. A film featuring O. brassicae is<br />

also available (Webster, 2006a).<br />

Epidermal cells and root hairs of infected<br />

cabbage roots contain one or more spherical or<br />

cylindrical thalli, sometimes filling the whole<br />

cell (Fig. 6.14a). The cy<strong>to</strong>plasm of the thallus<br />

is granular and the entire contents divide in<strong>to</strong><br />

numerous posteriorly uniflagellate zoospores<br />

that escape through one or more discharge<br />

tubes which penetrate the outer wall of the<br />

host cell (Temmink & Campbell, 1968). Release of<br />

the zoospores takes place within a few minutes<br />

of washing the roots free from soil. The tip of<br />

the discharge tube breaks down and zoospores<br />

rush out and swim actively in the water. The<br />

zoospores are very small, tadpole-like, with<br />

Fig 6.14 Olpidium brassicae in cabbage<br />

roots. (a) Two ripe sporangia and one<br />

empty sporangium in an epidermal cell.<br />

Each sporangium has a single exit tube.<br />

(b) Empty sporangium showing three exit<br />

tubes. (c) Zoospores. (d) Zoospore cysts<br />

on a root hair. Note that some cysts are<br />

uninucleate and some are binucleate.<br />

(e) Resting sporangia. (a,b,d,e) <strong>to</strong> same<br />

scale.


SPIZELLOMYCETALES<br />

147<br />

Fig 6.15 Olpidium brassicae.<br />

Diagrammatic representation of<br />

L.S. of zoospore (afterTemmink &<br />

Campbell,1969a).<br />

a spherical head and a long trailing flagellum.<br />

The fine structure of the zoospore is summarized<br />

in Fig 6.15. A distinctive feature is the banded<br />

rhizoplast which connects the kine<strong>to</strong>some <strong>to</strong><br />

the nucleus (Temmink & Campbell, 1969a;<br />

Lange & Olson, 1976a,b; Barr & Hartmann,<br />

1977). This structure has also been reported<br />

from the zoospore of the eucarpic chytrid<br />

Rhizophlyctis rosea (see p. 148; Barr & Hartmann,<br />

1977).<br />

The zoospores swim actively in water for<br />

about 20 min. If roots of cabbage seedlings are<br />

placed in a suspension of zoospores, these settle<br />

on the root hairs and epidermal cells, withdraw<br />

their flagella and encyst. The cysts are attached<br />

by a slime-like adhesive (Temmink & Campbell,<br />

1969b). The cyst wall and the root cell wall at<br />

the point of attachment are dissolved and<br />

the root cell is penetrated. The cyst contents<br />

are transferred <strong>to</strong> the inside of the host cell,<br />

probably by the enlargement of a vacuole which<br />

develops inside the cyst, whilst the empty cyst<br />

remains attached <strong>to</strong> the outside. The process<br />

of penetration can take place in less than<br />

one hour (Aist & Israel, 1977). Within 2 days of<br />

infection, small spherical thalli can be seen in<br />

the root hairs and epidermal cells of the root,<br />

carried around the cell by cy<strong>to</strong>plasmic streaming.<br />

The thalli enlarge and become multinucleate.<br />

Within 4 5 days discharge tubes develop<br />

and the thalli are ready <strong>to</strong> release zoospores.<br />

In some infected roots, stellate bodies with<br />

thick folded walls, lacking discharge tubes,<br />

are also found (Fig. 6.14e). These are resting<br />

sporangia. There is no evidence that they are<br />

formed as a result of sexual fusion either in<br />

O. brassicae or in O. bornovanus (Barr, 1988).<br />

Although biflagellate zoospores may occur in<br />

O. brassicae, these possibly result from incomplete<br />

cleavage (Temmink & Campbell, 1968) and<br />

zoospores with as many as 6 flagella have been<br />

observed (Garrett & Tomlinson, 1967). The resting<br />

sporangia are capable of germination<br />

7 10 days after they mature, and germinate by<br />

the formation of one or two exit papillae<br />

through which the zoospores escape.<br />

Virus transmission by Olpidium<br />

Several plant viruses are transmitted by zoospores<br />

of Olpidium. By analogy with plant virus<br />

transmission by aphids, Adams (1991) arbitrarily


148 CHYTRIDIOMYCOTA<br />

distinguished viruses with non-persistent and<br />

persistent transmission by fungi, although<br />

Campbell (1996) objected <strong>to</strong> the use of these<br />

terms, distinguishing instead between viruses<br />

which can be acquired in vitro (i.e. outside the<br />

plant) and those that can only be acquired in vivo<br />

(within the host cell).<br />

Tobacco necrosis virus (TNV) and cucumber<br />

necrosis virus (CNV) are non-persistent viruses<br />

which can be acquired in vitro by zoospores<br />

of O. brassicae or O. bornovanus (respectively).<br />

Virus particles (virions) are adsorbed on<strong>to</strong> the<br />

plasmalemma of the zoospore and on<strong>to</strong> the<br />

flagellar axonemal sheath which is continuous<br />

with it (Temmink et al., 1970). Binding seems <strong>to</strong><br />

occur between the virus coat and specific molecules<br />

at the zoospore surface, possibly oligosaccharide<br />

side chains of proteins (Kakani et al.,<br />

2003; Rochon et al., 2004). When the flagellum is<br />

withdrawn in<strong>to</strong> the body of the zoospore at<br />

encystment, virus particles are introduced in<strong>to</strong><br />

the fungal cy<strong>to</strong>plasm and are then transmitted<br />

in<strong>to</strong> the plant upon infection. Air-dried roots<br />

containing TNV virus and O. brassicae resting<br />

sporangia, or living virus-infected roots with<br />

resting sporangia treated with 5N HCl, were<br />

incapable of transmitting virus even though<br />

the resting sporangia survived these treatments,<br />

indicating that TNV is not carried inside the<br />

resting sporangia (Campbell & Fry, 1966).<br />

Lettuce big vein virus, LBVV, in contrast, is<br />

an example of the persistent type (Grogan<br />

et al., 1958). In this case it has been shown that<br />

the virus can persist in air-dried resting sporangia<br />

for 18 20 years (Campbell, 1985). Here<br />

the virions are acquired in vivo and they are<br />

present inside the zoospores which emerge from<br />

sporangia and resting sporangia (Campbell,<br />

1996).<br />

Classification of Olpidium<br />

Although previously classified within the<br />

family Olpidiaceae in the order Chytridiales,<br />

D. J. S. Barr (2001) has placed Olpidium in the<br />

order Spizellomycetales along with Rhizophlyctis<br />

on the basis of similarities in zoospore structure.<br />

Ribosomal DNA sequence comparisons are<br />

inconclusive in that they do not show any close<br />

similarity between Olpidium and either Chytridium<br />

or Spizellomyces (Ward & Adams, 1998).<br />

6.3.2 Rhizophlyctis<br />

There are about 10 known species of Rhizophlyctis<br />

with monocentric eucarpic thalli, growing as<br />

saprotrophs on a variety of substrata in soil,<br />

freshwater and the sea. Rhizophlyctis rosea grows<br />

on cellulose-rich substrata in soil, and it probably<br />

plays an active but currently underestimated<br />

role in cellulose decay (Powell, 1993). It can<br />

survive for prolonged periods in dry soil, even<br />

when this is heated <strong>to</strong> 90°C for two days (Gleason<br />

et al., 2004) and, in fact, the recovery of R. rosea<br />

is greatly enhanced if soil samples are air-dried<br />

prior <strong>to</strong> isolation experiments (Willoughby,<br />

2001). Willoughby (1998b) has estimated that<br />

over 1000 thallus-forming units could be recovered<br />

per gram of air-dry soil or leaf humus<br />

fragments from Provence, France. These numbers<br />

may arise from one or a few sporangia, since<br />

a single sporangium about 100 mm in diameter<br />

may discharge up <strong>to</strong> 30 000 zoospores. Mitchell<br />

and Deacon (1986) have shown that zoospores<br />

of R. rosea accumulate preferentially on cellulosic<br />

materials.<br />

The fungus is readily isolated and grown<br />

in culture, and details of techniques have<br />

been provided by Stanier (1942), Barr (1987),<br />

Willoughby (1998b) and Weber and Webster<br />

(2000a). The placing of a small crumb of soil<br />

on<strong>to</strong> moist tissue paper or cellophane overlying<br />

agar containing mineral salts, or the floating<br />

of squares of cellophane on water containing<br />

a soil sample, are followed within a few days<br />

by the development of thalli with bright pink<br />

sporangia. The sporangia are attached <strong>to</strong> coarse<br />

rhizoids which arise at several points on the<br />

sporangial wall and extend throughout the<br />

cellulosic substratum, tapering <strong>to</strong> fine points.<br />

Extensive corrosion of the substrate underneath<br />

the thallus and rhizoids points at the secretion<br />

of powerful cellulases (Fig. 6.17).<br />

Although the fungus is usually monocentric,<br />

there are also records of some polycentric<br />

isolates. When ripe, the sporangia have pink<br />

granular contents which differentiate in<strong>to</strong><br />

numerous uninucleate posteriorly uniflagellate


SPIZELLOMYCETALES<br />

149<br />

zoospores (Fig. 6.16a). One <strong>to</strong> several discharge<br />

tubes are formed, and the tip of each tube<br />

contains a clear mucilaginous plug which, prior<br />

<strong>to</strong> discharge, is exuded in a mass from the tip of<br />

the tube (Fig. 6.16c). While the plug of mucilage<br />

dissolves, the zoospores within the sporangium<br />

show active movement and then escape by<br />

swimming through the tube. In some specimens<br />

of R. rosea it has been found that a membrane<br />

may form over the cy<strong>to</strong>plasm at the base of<br />

the discharge tubes. If the sporangia do not<br />

discharge their spores immediately, the membrane<br />

may thicken. When spore discharge<br />

occurs, these thickened membranes can be seen<br />

floating free within the sporangia, and the term<br />

endo-operculum has been applied <strong>to</strong> them. The<br />

genus Karlingia was erected for forms possessing<br />

such endo-opercula, including R. rosea, which<br />

is therefore sometimes referred <strong>to</strong> as Karlingia<br />

rosea, but the validity of this separation is<br />

questionable because the presence or absence<br />

of endo-opercula is a variable character<br />

(Blackwell & Powell, 1999).<br />

Zoospores of R. rosea are capable of swimming<br />

for several hours. The head of the zoospore is<br />

often globose, but can become pear-shaped or<br />

show amoeboid changes in shape. It contains a<br />

prominent lipid body, several bright refringent<br />

globules, and bears a single trailing flagellum.<br />

Ultrastructural details resemble those of<br />

Olpidium brassicae in the presence of a striated<br />

rhizoplast connecting kine<strong>to</strong>some and nucleus<br />

(Barr & Hartmann, 1977). On coming <strong>to</strong> rest on<br />

a suitable substratum, the flagellum is withdrawn<br />

and the body of the zoospore enlarges<br />

<strong>to</strong> form the rudiment of the sporangium, whilst<br />

rhizoids appear at various points on its surface.<br />

Within the sporangium, the flagella are tightly<br />

wrapped around the zoospores (Chambers<br />

& Willoughby, 1964).<br />

Resting sporangia are also found. They are<br />

brown, globose or angular and have a thickened<br />

Fig 6.16 Rhizophlyctisrosea. (a) Zoospores. (b) Young thallus formed on germination of zoospore.The zoospore cyst has enlarged<br />

and will form the sporangium. (c) Older sporangium showing three discharge tubes. (d) Sporangium showing mucilage plugs at the<br />

tips of the discharge tubes and thickenings of the cell membrane at the bases of the tubes. Such thickenings are termed<br />

endo-opercula. (e) Globose sporangium and seven visible papillae. (f) Resting sporangium formed inside an empty zoosporangium.<br />

(a,b) <strong>to</strong> same scale; (c f) <strong>to</strong> same scale.


150 CHYTRIDIOMYCOTA<br />

Fig 6.17 Scanning electron<br />

micrograph of two thalli of<br />

Rhizophlyctisrosea on a cellophane<br />

membrane. Pit corrosion is visible<br />

where a thallus has been lifted<br />

from the substratum (arrows).<br />

wall (Fig. 6.16f ). Whether they are formed<br />

sexually in R. rosea is not known. Couch (1939)<br />

has, however, put forward evidence that the<br />

fungus is heterothallic because single isolates<br />

grown in culture failed <strong>to</strong> produce resting<br />

sporangia whereas these structures did form<br />

when certain cultures were paired. Stanier (1942)<br />

has reported the occurrence of biflagellate zoospores,<br />

but whether these represented zygotes<br />

seemed doubtful. In the homothallic chitinophilic<br />

fungus Rhizophlyctis oceanis, Karling (1969)<br />

has described frequent fusions between zoospores.<br />

These fusions are possibly sexual, but<br />

unfortunately Karling was unable <strong>to</strong> cultivate<br />

the resulting thalli <strong>to</strong> the stage of resting spore<br />

development.<br />

On germination, the resting sporangium of<br />

R. rosea functions as a prosporangium, although<br />

it is uncertain whether resting sporangia are<br />

important for survival in nature. Willoughby<br />

(2001) has shown that R. rosea could be recovered<br />

from cellophane baits in as little as 5 6 h after<br />

placing air-dried soil samples in water, and it<br />

was concluded that these zoospores were derived<br />

from sporangia instead of resting spores which<br />

need a longer time <strong>to</strong> produce zoospores.<br />

The nutritional requirements of R. rosea are<br />

simple. It shows vigorous growth on cellulose<br />

as the sole carbon source but it can utilize<br />

a range of carbohydrates such as glucose,<br />

cellobiose and starch. The pink colour of the<br />

sporangia is due <strong>to</strong> the presence of carotenoid<br />

pigments such as g-carotene, lycopene and a<br />

xanthophyll.<br />

6.4 Neocallimastigales<br />

(rumen fungi)<br />

A very interesting and unusual group of zoosporic<br />

fungi inhabits the rumens (foreguts) of<br />

ruminants (herbivorous mammals which regurgitate<br />

and masticate previously ingested food)<br />

like cows, sheep and deer. They have also been<br />

found in some non-ruminants such as horses and<br />

elephants and probably occur in the guts of<br />

many large herbivores. These fungi are obligate<br />

anaerobes which can flourish in the rumen<br />

because oxygen is depleted there by the intense<br />

respira<strong>to</strong>ry activity of a dense population of<br />

pro<strong>to</strong>zoa and bacteria, some of which are<br />

facultative anaerobes capable of scavenging free<br />

oxygen. Their zoospores were at first thought <strong>to</strong><br />

be pro<strong>to</strong>zoa and were not recognized as belonging<br />

<strong>to</strong> fungi because obligately anaerobic fungi


NEOCALLIMASTIGALES (RUMEN FUNGI)<br />

151<br />

were not believed <strong>to</strong> exist. Further, microbiologists<br />

working on microbes from the ruminant<br />

gut studied only strained rumen fluid and<br />

therefore failed <strong>to</strong> see the thalli of fungi attached<br />

<strong>to</strong> herbage fragments. The view that the motile<br />

cells swimming in rumen fluid belonged <strong>to</strong><br />

flagellates was challenged by Orpin (1974), who<br />

observed that there was an enormous increase in<br />

the concentration of ‘flagellates’ in the rumen of<br />

sheep within a short time of feeding. The ratio of<br />

minimum (pre-feeding) <strong>to</strong> maximum concentration<br />

of motile cells could vary between 1:15 and<br />

1:296 (average 1:47), and if these were organisms<br />

reproducing by binary fission it would be<br />

necessary for them <strong>to</strong> undergo six successive<br />

cell divisions in 15 min. The explanation for the<br />

rapid increase in motile cells is that sedentary<br />

fungal thalli, anchored by rhizoids <strong>to</strong> partially<br />

digested food fragments floating in the rumen,<br />

are stimulated <strong>to</strong> release zoospores by soluble<br />

substances such as haems released from the<br />

newly ingested food material. The zoospores<br />

attach themselves in large numbers <strong>to</strong> the<br />

herbage fragments, and germinate <strong>to</strong> form<br />

rhizoidal or rhizomycelial thalli with sporangia<br />

capable of releasing further zoospores within<br />

about 30 h.<br />

Some 5 genera and 15 species have now been<br />

distinguished (Theodorou et al., 1992, 1996;<br />

Trinci et al., 1994). They include Caecomyces<br />

which has mono- and polycentric thalli,<br />

Anaeromyces and Orpinomyces with polycentric<br />

thalli, and Piromyces and Neocallimastix which<br />

are monocentric. The zoospores of Anaeromyces,<br />

Caecomyces and Piromyces are uniflagellate<br />

whilst those of Neocallimastix and Orpinomyces<br />

are multiflagellate (see Fig. 6.18). They were<br />

classified within the order Spizellomycetales,<br />

family Callimasticaceae by Heath et al. (1983)<br />

and Barr et al. (1989) but are now placed in a<br />

separate order Neocallimastigales (Li et al., 1993;<br />

D. J. S. Barr, 2001). Special techniques and media<br />

are needed for isolating and handling anaerobic<br />

fungi, but the life cycle details of several have<br />

now been followed in pure culture. One of the<br />

best known is N. hurleyensis, isolated from sheep<br />

(Fig. 6.18). Minutes after the arrival of fresh<br />

food, globose ripe zoosporangia on previously<br />

colonized grass fragments release zoospores<br />

through an apical pore and these attach themselves<br />

<strong>to</strong> herbage fragments and germinate <strong>to</strong><br />

produce rhizoids which penetrate and digest the<br />

ingested plant material. The walls of the thallus<br />

contain chitin. A single zoosporangium develops<br />

and is cut off from the rhizoidal system by<br />

a septum. The rhizoidal part is devoid of nuclei,<br />

but within the zoosporangium repeated nuclear<br />

divisions occur before the cy<strong>to</strong>plasm cleaves<br />

<strong>to</strong> form 64 128 zoospores. The life cycle of<br />

N. hurleyensis from zoospore germination <strong>to</strong> the<br />

release of a fresh crop of zoospores lasts about<br />

29 31 h at 39°C (Lowe et al., 1987a). The<br />

zoospores bear 8 16 whiplash flagella inserted<br />

posteriorly in two rows.<br />

The ultrastructure of zoospores has been<br />

described for several species of Neocallimastix,<br />

including N. patriciarum (Orpin & Munn, 1986),<br />

N. frontalis (Munn et al., 1981; Heath et al., 1983)<br />

and N. hurleyensis (Webb & Theodorou, 1988,<br />

1991). There are differences in detail. For<br />

example, the zoospore of N. frontalis has a waistlike<br />

constriction, with the majority of the cy<strong>to</strong>plasmic<br />

organelles concentrated in the posterior<br />

portion near the insertion of the flagella.<br />

Characteristic organelles known from zoospores<br />

of aerobic chytridiomycetes such as mi<strong>to</strong>chondria,<br />

Golgi bodies, lipid droplets or gamma<br />

particles (seen in zoospores of Blas<strong>to</strong>cladiella<br />

emersonii; Fig. 6.19) are absent. In the posterior<br />

portion of the zoospore of N. hurleyensis near the<br />

point of insertion of the flagella, an irregularly<br />

shaped complex structure interpreted as a<br />

hydrogenosome has been reported in place of a<br />

mi<strong>to</strong>chondrion. In zoospores of N. patriciarum<br />

there are many presumed hydrogenosomes<br />

concentrated around the region of flagellar<br />

insertion. Hydrogenosomes are organelles<br />

capable of the anaerobic metabolism of hexoses<br />

<strong>to</strong> acetic and formic acids. Pro<strong>to</strong>ns (H þ ) act as<br />

electron accep<strong>to</strong>rs, so that gaseous H 2 is released<br />

by the activity of the enzyme hydrogenase<br />

(Müller, 1993; Boxma et al., 2004). The hydrogen,<br />

in turn, is used by anaerobic methanogenic<br />

bacteria <strong>to</strong> reduce CO 2 <strong>to</strong> CH 4 (methane) which<br />

escapes in profusion through the front and hind<br />

exits of the ruminant digestive tracts.<br />

Hydrogenosomes are found in several anaerobic<br />

lower eukaryotes and are believed <strong>to</strong> be derived


152 CHYTRIDIOMYCOTA<br />

Fig 6.18 Neocallimastix hurleyensis. (a) Rhizoidal<br />

thallus with zoosporangium. (b) Release of<br />

zoospores. (c) Tracing of T.E.M. of zoospore<br />

with 14 flagella in longitudinal section.<br />

Diagrams based on Webb and<br />

Theodorou (1991).<br />

from mi<strong>to</strong>chondria (Embley et al., 2002). Whereas<br />

mi<strong>to</strong>chondria of most fungi contain a limited<br />

amount of DNA, hydrogenosomes of rumen<br />

chytrids seem <strong>to</strong> have lost their genome al<strong>to</strong>gether<br />

(Bullerwell & Lang, 2005).<br />

Granular inclusion bodies which contain<br />

aggregates of ribosome-like particles and also<br />

free ribosome-like arrays are found anterior <strong>to</strong><br />

the nucleus. Rosettes of glycogen represent the<br />

energy reserve of the zoospore. The shafts of<br />

the flagella contain the familiar eukaryotic 9 þ 2<br />

arrangement of microtubules, but in N. frontalis<br />

the microtubules do not extend in<strong>to</strong> the tips<br />

of the flagella which are narrower than the<br />

proximal part.<br />

Ecologically, these anaerobic fungi play an<br />

important role in the early colonization of<br />

ingested herbage and have a wide range of<br />

enzymes which enable them <strong>to</strong> utilize monosaccharides,<br />

disaccharides and polysaccharides<br />

such as xylan, cellulose, starch and glycogen<br />

(Theodorou et al., 1992). They may play an active<br />

role in fibre breakdown. It is likely that colonization<br />

of straw particles by these fungi aids further<br />

attack by bacteria. The survival of anaerobic<br />

fungi outside the unusual and protective environment<br />

of the herbivore gut occurs in dried<br />

faeces in the form of cysts or as melanized<br />

thick-walled thalli whilst transmission <strong>to</strong> young<br />

animals takes place in saliva during licking


BLASTOCLADIALES<br />

153<br />

and grooming (Lowe et al., 1987b; Wubah et al.,<br />

1991; Theodorou et al., 1992).<br />

6.5 Blas<strong>to</strong>cladiales<br />

6.5.1 <strong>Introduction</strong><br />

Species belonging <strong>to</strong> the Blas<strong>to</strong>cladiales are<br />

mostly saprotrophs in soil, water, mud or aquatic<br />

plant and animal debris, and some are pathogens<br />

of plants, invertebrate animals or fungi. Most<br />

are obligate aerobes, but Blas<strong>to</strong>cladia spp. are<br />

facultatively anaerobic, requiring a fermentable<br />

substrate and growing on submerged fleshy<br />

fruits, twigs or other plant materials rich in<br />

soluble carbohydrates (Emerson & Robertson,<br />

1974). The life cycles of Blas<strong>to</strong>cladiales show<br />

great variations and in some forms there is an<br />

alternation of distinct haploid game<strong>to</strong>thallic<br />

and diploid sporothallic generations. These<br />

terms are used in preference <strong>to</strong> the botanical<br />

terms game<strong>to</strong>phytic and sporophytic. Species<br />

of Physoderma, previously grouped with the<br />

Chytridiales (Lange & Olson, 1980), are biotrophic<br />

parasites of higher plants (Karling, 1950). They<br />

include P. maydis, the cause of brown spot of<br />

maize, and P. alfalfae (Lange et al., 1987). One<br />

genus, Coelomomyces, consists of obligate parasites<br />

of insects, usually mosqui<strong>to</strong> larvae (Couch &<br />

Bland, 1985). This genus is unusual in that<br />

the vegetative thallus is a wall-less plasmodiumlike<br />

structure lacking rhizoids. The life cycle<br />

is completed in unrelated alternate animal<br />

hosts, sporothalli occurring in mosqui<strong>to</strong> larvae<br />

(Insecta) and game<strong>to</strong>thalli in a copepod<br />

(Crustacea) (Whisler et al., 1975; Federici, 1977).<br />

Attempts are being made <strong>to</strong> use Coelomomyces<br />

in the biological control of mosqui<strong>to</strong>es. Catenaria<br />

anguillulae, a facultative parasite of nema<strong>to</strong>des<br />

and their eggs, liver fluke eggs and some other<br />

invertebrates, can be grown in culture (Couch,<br />

1945; Barron, 1977; Bars<strong>to</strong>w, 1987), whilst<br />

Catenaria allomycis is a biotrophic parasite of<br />

Allomyces (Couch, 1945; Sykes & Porter, 1980).<br />

With the exception of Coelomomyces, the<br />

thallus of members of the Blas<strong>to</strong>cladiales is<br />

eucarpic. The morphologically simpler forms<br />

such as Blas<strong>to</strong>cladiella (Fig. 6.22) are monocentric,<br />

with a spherical or sac-like zoosporangium or<br />

resting sporangium arising directly or on<br />

a short one-celled stalk from a tuft of radiating<br />

rhizoids. These simpler types show considerable<br />

similarity <strong>to</strong> monocentric Chytridiales of<br />

other orders such as Rhizophlyctis rosea (Figs. 6.16<br />

and 6.17), and in the vegetative state they<br />

may be difficult <strong>to</strong> distinguish. The more<br />

complex organisms such as Allomyces are polycentric,<br />

and the thallus is differentiated in<strong>to</strong><br />

a trunk-like portion which has rhizoids below<br />

whilst branching above, often dicho<strong>to</strong>mously,<br />

and bearing sporangia of various kinds at the<br />

tips of the branches. Chitin has been demonstrated<br />

in the walls of Allomyces and Blas<strong>to</strong>cladiella<br />

(Porter & Jaworski, 1966; Youatt, 1977; Maia,<br />

1994).<br />

The zoospore of Blas<strong>to</strong>cladiales<br />

The zoospore of Blas<strong>to</strong>cladiales has a single<br />

posterior flagellum of the whiplash type.<br />

Details of the fine structure of this kind of<br />

zoospore have been reviewed by Fuller (1976)<br />

and Lange and Olson (1979). The best known<br />

are Blas<strong>to</strong>cladiella emersonii (Cantino et al., 1963;<br />

Reichle & Fuller, 1967) and Allomyces macrogynus<br />

(Fuller & Olson, 1971). The structure of the<br />

zoospore of B. emersonii is summarized diagrammatically<br />

in Fig. 6.19. The zoospore is tadpolelike<br />

with a pear-shaped head about 7 9 mm and<br />

a single, trailing flagellum about 20 mm long.<br />

Under the light microscope, the most conspicuous<br />

internal structure is the dense crescentshaped<br />

nuclear cap which surrounds the more<br />

transparent nucleus. The nuclear cap is rich in<br />

RNA and protein, and is filled with ribosomes.<br />

The zoospore of B. emersonii is unusual in that<br />

it contains only a single large mi<strong>to</strong>chondrion,<br />

situated near the flagellar kine<strong>to</strong>some.<br />

The organization of the flagellum is essentially<br />

as described on p. 129. The nine triplet<br />

microtubules extend in a funnel-shaped manner<br />

from the proximal end of the kine<strong>to</strong>some<br />

<strong>to</strong>wards the nucleus and nuclear cap, maintaining<br />

its conical shape. Extending in<strong>to</strong> the mi<strong>to</strong>chondrion<br />

and linking up the kine<strong>to</strong>some


154 CHYTRIDIOMYCOTA<br />

Fig 6.19 Blas<strong>to</strong>cladiella emersonii zoospore,<br />

fine structure, diagrammatic and not <strong>to</strong> scale.<br />

(a) L.S. of zoospore along the axis of the<br />

flagellum. (b) T.S. of kine<strong>to</strong>some showing nine<br />

triplets of microtubules. (c) T.S. of kine<strong>to</strong>some<br />

at a slightly lower level showing the origin of<br />

two of the banded rootlets which extend<br />

in<strong>to</strong> the mi<strong>to</strong>chondrion.The cristae of the<br />

mi<strong>to</strong>chondrion are close <strong>to</strong> the membrane<br />

which surrounds the banded rootlets.<br />

(d) T.S. of axoneme showing the nine paired<br />

peripheral microtubules and the two central<br />

microtubules.<br />

with it are three striated bodies variously<br />

referred <strong>to</strong> as flagellar rootlets, striated rootlets<br />

or banded rootlets. They are contained within<br />

separate channels, and each is surrounded by<br />

a unit membrane. Since the energy for propulsion<br />

is generated within the mi<strong>to</strong>chondrion, it is<br />

possible that the banded rootlets are, in some<br />

way, responsible for transmitting energy <strong>to</strong> the<br />

base of the axoneme. It is also possible that the<br />

banded rootlets serve <strong>to</strong> anchor the flagellum<br />

within the body of the zoospore.<br />

There are two other obvious kinds of organelle<br />

within the body of the Blas<strong>to</strong>cladiella zoospore.<br />

The lipid sac attached <strong>to</strong> the mi<strong>to</strong>chondrion<br />

contains a group of lipid droplets which is<br />

surrounded by a unit membrane. It is not<br />

known whether lipid forms the energy reserve<br />

used in swimming, cy<strong>to</strong>plasmic glycogen deposits<br />

being a more plausible alternative (Cantino<br />

et al., 1968). In the anterior of the zoospore<br />

between the nuclear cap and the plasma<br />

membrane, there is a group of granules about<br />

0.5 mm in diameter, called gamma particles.<br />

They consist of an inner core, shaped like<br />

an elongated cup and bearing two unequal<br />

openings at opposite sides of the cup. This cupshaped<br />

structure is enveloped in a unit membrane<br />

(Myers & Cantino, 1974). Gamma particles<br />

are only present in developing and motile<br />

zoospores but disappear as the zoospore encysts.<br />

Formerly thought <strong>to</strong> represent the chytrid<br />

equivalent of the chi<strong>to</strong>some found in higher<br />

fungi (see p. 6), this notion has now been<br />

discarded (Hohn et al., 1984).


BLASTOCLADIALES<br />

155<br />

The zoospore of Allomyces macrogynus broadly<br />

resembles that of B. emersonii (Fuller & Olson,<br />

1971). Gamma particles are present in the<br />

zoospore and, during encystment, these form<br />

vesicles which fuse with the plasma membrane.<br />

Fusion coincides with the appearance of<br />

wall material around the cyst (Bars<strong>to</strong>w &<br />

Pommerville, 1980). The zoospore of Allomyces<br />

differs from that of Blas<strong>to</strong>cladiella in some other<br />

ways. Although there is a large basal mi<strong>to</strong>chondrion,<br />

many smaller mi<strong>to</strong>chondria are also<br />

present, generally located along the membrane<br />

of the nuclear cap in the anterior part of the<br />

cell. A complex structure situated laterally at<br />

the base of the body of the zoospore, between<br />

the nucleus and the zoospore membrane, has<br />

been termed the side body complex by Fuller<br />

and Olson (1971). It consists of two closely<br />

appressed membranes separated by an electronopaque<br />

material. These membranes subtend<br />

numerous electron-opaque, membrane-bound<br />

bodies, lipid bodies and a portion of the basal<br />

mi<strong>to</strong>chondrion. In addition, there are membrane-bound<br />

non-lipid bodies termed Stüben<br />

bodies by Fuller and Olson (1971), whose function<br />

and composition are uncertain.<br />

The zoospore is propelled forward by rhythmic<br />

lashing of the flagellum, and it can swim<br />

for a period even under anaerobic conditions.<br />

It is also capable of amoeboid changes of shape.<br />

On coming <strong>to</strong> rest, the flagellum is retracted<br />

in<strong>to</strong> the body of the zoospore. There are different<br />

interpretations of the manner in which flagellar<br />

retraction is achieved. Cantino et al. (1968)<br />

have suggested that the flagellum is retracted<br />

by a revolving action of the nucleus, whereas<br />

in the ‘lash-around’ mechanism the flagellum<br />

coils around the body of the spore, the flagellar<br />

membrane fuses with the plasmalemma of<br />

the spore and the axoneme enters the spore<br />

cy<strong>to</strong>plasm (Olson, 1984). In Allomyces, the zoospore<br />

cyst produces, at one point, a narrow germtube<br />

which branches <strong>to</strong> form the rhizoidal<br />

system. At the opposite pole, the zoospore cyst<br />

forms a wider germ tube which gives rise <strong>to</strong><br />

hyphae which branch and later bear sporangia.<br />

This bipolar germination pattern is a point<br />

of difference between the Blas<strong>to</strong>cladiales and<br />

the Chytridiales, in which germination is<br />

typically unipolar. The rhizoids are strongly<br />

chemotropic and specialize in nutrient uptake<br />

and transport. An inwardly directed electrical<br />

current has been detected around the rhizoids,<br />

and an outwardly directed current around<br />

the hyphae and hyphal tips. The inward current<br />

at the rhizoids may be the consequence of<br />

localized pro<strong>to</strong>n-driven solute transport<br />

(de Silva et al., 1992).<br />

Life cycles of Blas<strong>to</strong>cladiales<br />

A number of distinct life his<strong>to</strong>ry patterns are<br />

found. In Allomyces arbuscula, for example, isomorphic<br />

alternation of haploid game<strong>to</strong>thallic<br />

and diploid sporothallic generations has been<br />

demonstrated. In A. neo-moniliformis (¼ A. cys<strong>to</strong>genes)<br />

there is no free-living sexual generation,<br />

but this stage is represented by a cyst (see<br />

below). In A. anomalus, only the asexual stage<br />

has been found in normal cultures, but experimental<br />

treatments may result in the development<br />

of sexual thalli. Similar variations in life<br />

cycles have been found in other genera such<br />

as Blas<strong>to</strong>cladiella. A characteristic feature of<br />

the asexual thalli of the Blas<strong>to</strong>cladiales is the<br />

presence of resting sporangia with chitinous,<br />

pitted walls impregnated with a dark brown,<br />

melanin-type pigment. The pits are inwardly<br />

directed conical pores in the wall. The inner<br />

ends of the pores abut against a smooth, colourless<br />

inner layer of wall material surrounding<br />

the cy<strong>to</strong>plasm (Skucas, 1967, 1968). The resting<br />

sporangia of Allomyces can remain viable for up<br />

<strong>to</strong> 30 years in dried soil. The ease with which<br />

certain members of the group can be grown<br />

in culture has facilitated extensive studies of<br />

their nutrition and physiology, and the results<br />

of some of these investigations are discussed<br />

below.<br />

Four families have been recognized<br />

Coelomomycetaceae, Catenariaceae, Physodermataceae<br />

and Blas<strong>to</strong>cladiaceae but of these we<br />

shall study only Allomyces and Blas<strong>to</strong>cladiella,<br />

both representatives of the Blas<strong>to</strong>cladiaceae.<br />

6.5.2 Allomyces<br />

Species of Allomyces are found in mud or soil of<br />

the tropics or subtropics, including desert soil,


156 CHYTRIDIOMYCOTA<br />

and if dried samples of soil are placed in water<br />

and ‘baited’ with boiled hemp seeds, the baits<br />

may become colonized by zoospores. From such<br />

material, it is possible <strong>to</strong> obtain pure cultures by<br />

streaking or pipetting zoospores on<strong>to</strong> suitable<br />

agar media and <strong>to</strong> follow the complete life<br />

his<strong>to</strong>ry of these fungi in the labora<strong>to</strong>ry. Olson<br />

(1984) has given a full account of the taxonomy,<br />

life cycles, morphogenesis and genetics of<br />

different species of Allomyces, with practical<br />

details of how <strong>to</strong> grow and handle them. Good<br />

growth occurs on a medium containing yeast<br />

extract, pep<strong>to</strong>ne and soluble starch (YPsS), but<br />

chemically defined media have also been used.<br />

There is a requirement for thiamine and organic<br />

nitrogen in the form of amino acids.<br />

Emerson (1941) isolated species of Allomyces<br />

from soil samples from all over the world. He<br />

distinguished three types of life his<strong>to</strong>ry, represented<br />

by three subgenera.<br />

Sub-genus Eu-Allomyces<br />

The Eu-Allomyces type of life his<strong>to</strong>ry is exemplified<br />

by A. arbuscula and A. macrogynus (Fig. 6.21; for<br />

a film, see Webster & Hard, 1998a). Resting<br />

sporangia are formed on asexual diploid thalli.<br />

They contain about 12 nuclei which undergo<br />

meiosis during the early stages of germination<br />

(Olson, 1974). The cy<strong>to</strong>plasm cleaves around the<br />

48 haploid nuclei <strong>to</strong> form the zoospores. Since<br />

meiosis occurs in the resting sporangia, these<br />

have been termed meiosporangia, and the<br />

haploid zoospores meiospores. The meiospores<br />

are released when the outer wall of the brown<br />

pitted resting sporangium cracks open by a slit<br />

and the inner wall balloons outwards and<br />

eventually opens by one or more pores. The<br />

meiospores swim by movement of the trailing<br />

flagellum and, on coming <strong>to</strong> rest, encyst and<br />

germinate as described above <strong>to</strong> form a rhizoidal<br />

system and a trunk-like region which bears<br />

dicho<strong>to</strong>mous branches.<br />

The tips of the branches have been claimed<br />

<strong>to</strong> resemble the Spitzenkörper of higher fungi<br />

in being actin-rich, although secre<strong>to</strong>ry vesicles<br />

and/or microvesicles (chi<strong>to</strong>somes) have not been<br />

clearly shown (Srinivasan et al., 1996). Repeated<br />

nuclear division occurs <strong>to</strong> form a coenocytic<br />

structure, and finger-like ingrowths from the<br />

walls of the trunk-region and branches form<br />

incomplete septa, sometimes termed pseudosepta,<br />

with a pore in the centre through which<br />

cy<strong>to</strong>plasmic connections can be seen (Fig. 6.20d;<br />

Meyer & Fuller, 1985). The haploid thalli which<br />

develop from the meiospores are game<strong>to</strong>thallic,<br />

i.e. sexual. They are monoecious, and the tips<br />

of their branches swell <strong>to</strong> form paired sacs the<br />

male and female gametangia. The male gametangia<br />

can be identified by the presence of<br />

a bright orange pigment, g-carotene, whilst the<br />

female gametangia are colourless. In A. arbuscula<br />

the male gametangium is subterminal or<br />

hypogynous, i.e. beneath the terminal female<br />

gametangium, but in A. macrogynus the positions<br />

are reversed and the male gametangium is<br />

terminal or epigynous (Figs. 6.20e,i). The gametangia<br />

bear a number of colourless papillae on<br />

their walls, blocked by pulley-shaped plugs<br />

which eventually dissolve.<br />

The contents of the gametangia differentiate<br />

in<strong>to</strong> uninucleate gametes which differ in size<br />

and pigmentation. The female gametangium<br />

forms larger, colourless motile gametes (swarmers)<br />

whilst the male gametangium releases<br />

smaller, more active, orange-coloured swarmers.<br />

After escaping through the papillae in the walls<br />

of the gametangia, the gametes swim for a time<br />

and then pair off. A female gamete which fails<br />

<strong>to</strong> pair can function as a zoospore by germinating<br />

<strong>to</strong> form a new sexual thallus. A hormone,<br />

sirenin, is secreted by female gametangia during<br />

game<strong>to</strong>genesis and by the released female<br />

gametes, and this stimulates a chemotactic<br />

response in male gametes at the extremely low<br />

concentration of 8 10 11 M (Machlis, 1972;<br />

Carlile, 1996a). The chemical structure of sirenin<br />

has been determined (Fig. 6.22), and both d- and<br />

l-forms have been synthesized. Only l-sirenin is<br />

active. It is a bicyclic sesquiterpene, probably<br />

derived from the parent hydrocarbon sesquicarene<br />

(Nutting et al., 1968; Plattner & Rapoport,<br />

1971). A second hormone, parisin, which attracts<br />

female gametes, is secreted by male gametes.<br />

Its structure has not been determined, although<br />

it may well be related <strong>to</strong> sirenin (Pommerville<br />

& Olson, 1987).<br />

The biflagellate zygote resulting from the<br />

fusion of two gametes may swim for a while


BLASTOCLADIALES<br />

157<br />

Fig 6.20 (a h) Allomyces arbuscula. (a) Zoospores (haploid meiospores). (b) Young game<strong>to</strong>thalli, 24 h old. (c) Young sporothalli,18 h<br />

old. (d) Sporothallus, 30 h old. Perforations are visible in some of the septa. (e) Gametangia at the tips of the branches of the<br />

game<strong>to</strong>thallus. Note the disparity in the size of the gametes (anisogamy).The smaller male gametes are orange in colour whilst<br />

the larger female gametes are colourless.Compare the hypogynous arrangement of the male gametangia with the epigynous<br />

arrangement in A. macrogynus shown at (i). (f) Meiosporangia (resting sporangia, R.S.) and mi<strong>to</strong>sporangia (zoosporangia, Z.S.) on a<br />

sporothallus. (g) Release of mi<strong>to</strong>spores from zoosporangia (¼ mi<strong>to</strong>sporangia) on sporothallus. (h) Rupture of meiosporangium<br />

(¼ resting sporangium). (i) Allomyces macrogynus. Branch tip from game<strong>to</strong>thallus showing the arrangement of gametangia with<br />

terminal, epigynous male gametangia and anisogamous gametes.<br />

before it encysts and casts off the flagella.<br />

Nuclear fusion then follows (Pommerville &<br />

Fuller, 1976). The zygote develops immediately<br />

in<strong>to</strong> a diploid asexual thallus which differs<br />

from game<strong>to</strong>thalli in bearing two types of<br />

zoosporangia instead of gametangia. The first<br />

formed are thin-walled papillate zoosporangia<br />

formed singly or in rows at the tips of the


158 CHYTRIDIOMYCOTA<br />

Fig 6.21 Life cycle diagram of Eu-Allomyces as exemplified by A. macrogynus. A diploid sporothallus may produce diploid mi<strong>to</strong>spores<br />

from a colourless, thin-walled papillate mi<strong>to</strong>sporangium, and haploid meiospores from a thick-walled pitted meiosporangium in<br />

which meiosis occurs. Meiospores germinate <strong>to</strong> form a haploid gamethothallus which produces two different gametangia and<br />

releases haploid gametes of two kinds, small carotenoid-rich (shaded) ‘male’gametes and larger colourless ‘female’ones.Upon<br />

copulation, a diploid zygote gives rise <strong>to</strong> a sporothallus. Alternatively, if failing <strong>to</strong> pair up, the female gametes may function as<br />

zoospores, in which case they give rise <strong>to</strong> a new game<strong>to</strong>thallus. Small open circles represent haploid nuclei whereas diploid nuclei<br />

are drawn larger and split. It should be noted that many field strains of A. macrogynus have a higher ploidy level, e.g. alternating<br />

between diploid (small circles) and tetraploid (large split circles) conditions. Key events in the life cycle are plasmogamy (P),<br />

karyogamy (K) and meiosis (M).<br />

branches (Fig. 6.20g). Within these thin-walled<br />

sporangia the nuclei undergo mi<strong>to</strong>sis. Initially<br />

the nuclei are arranged in the cortical region<br />

of the cy<strong>to</strong>plasm, but later they migrate and<br />

become uniformly spaced apart. Movement of<br />

the nuclei is controlled by forces generated by<br />

actin microfilaments whilst their spacing and<br />

positioning is controlled by microtubules (Lowry<br />

et al., 1998). Cleavage of the cy<strong>to</strong>plasm around<br />

the nuclei <strong>to</strong> form diploid colourless zoospores<br />

is initiated by the formation of membranes seen<br />

first at the plasmalemma, then extending in<strong>to</strong><br />

the cortex <strong>to</strong> form a complex membranous<br />

network (Fisher et al., 2000). The process of<br />

cy<strong>to</strong>kinesis, i.e. the extension and fusion of<br />

Fig 6.22 Chemical structure of the hormone l-sirenin which<br />

attracts male gametes of Allomyces macrogynus.The structure<br />

of parisin, attractive <strong>to</strong> female gametes, does not seem <strong>to</strong> have<br />

been elucidated as yet.<br />

membranes, seems <strong>to</strong> be mediated principally<br />

by the actin component of the cy<strong>to</strong>skele<strong>to</strong>n<br />

(Lowry et al., 2004). According <strong>to</strong> Barron and<br />

Hill (1974), the development of the cleavage


BLASTOCLADIALES<br />

159<br />

membranes is induced by the availability of free<br />

water. Zoospores are released from the sporangia<br />

after dissolution of the plugs blocking the exit<br />

papillae. Since nuclear division in the thinwalled<br />

sporangia is mi<strong>to</strong>tic, these are termed<br />

mi<strong>to</strong>sporangia, and the diploid swarmers they<br />

release are mi<strong>to</strong>spores. The mi<strong>to</strong>spores, after a<br />

swimming phase, encyst and are capable of<br />

immediate germination, developing in<strong>to</strong> a<br />

further diploid asexual thallus.<br />

The second type of zoosporangium is the dark<br />

brown, thick-walled, pitted resting sporangium<br />

(meiosporangium), formed at the tips of the<br />

branches. Meiotic divisions within these sporangia<br />

result in the formation of the haploid<br />

meiospores, which develop in<strong>to</strong> sexual thalli.<br />

The life cycle of a member of the subgenus<br />

Eu-Allomyces is thus an isomorphic alternation<br />

of game<strong>to</strong>thallic and sporothallic generations<br />

(Fig. 6.21). Comparisons of the nutrition and<br />

physiology of the two generations show no<br />

essential distinction between them up <strong>to</strong> the<br />

point of production of gametangia or sporangia.<br />

Emerson and Wilson (1954) have made cy<strong>to</strong>logical<br />

and genetic studies of a number of<br />

collections of Allomyces. Interspecific hybrids<br />

between A. arbuscula and A. macrogynus have<br />

been produced in the labora<strong>to</strong>ry, and it has<br />

been shown that the fungus earlier described<br />

as A. javanicus is a naturally occurring hybrid<br />

between these two species. Cy<strong>to</strong>logical examination<br />

of the two parent species and of artificial<br />

and natural hybrids showed a great variation<br />

in chromosome number. In A. arbuscula the<br />

basic haploid chromosome number is 8, but<br />

strains with 16, 24 and 32 chromosomes have<br />

been found. In A. macrogynus the lowest haploid<br />

number encountered is 14, but strains with<br />

28 and 56 chromosomes are also known. The<br />

demonstration that these two species each<br />

represent a polyploid series was the first <strong>to</strong><br />

be made in fungi. The wild-type strain of<br />

A. macrogynus appears <strong>to</strong> be an au<strong>to</strong>tetraploid<br />

which, after meiosis, produces diploid game<strong>to</strong>thalli<br />

(Olson & Reichle, 1978).<br />

The behaviour of the hybrid strains is of<br />

considerable interest. As seen above, the parent<br />

species differ in the arrangement of the primary<br />

pairs of gametangia, A. arbuscula being<br />

hypogynous whilst A. macrogynus is epigynous.<br />

Following fusion of gametes derived from different<br />

parents, zygotes formed, germinated and<br />

gave rise <strong>to</strong> sporothalli. The meiospores from the<br />

hybrid sporothalli had a low viability (0.1 3.2%),<br />

as compared with a viability of about 63% for<br />

A. arbuscula meiospores, but some germinated<br />

<strong>to</strong> form game<strong>to</strong>thalli. The arrangement of the<br />

gametangia on these F 1 game<strong>to</strong>thalli showed<br />

a complete range from 100% epigyny <strong>to</strong> 100%<br />

hypogyny. Also, in certain game<strong>to</strong>thalli the ratio<br />

of male <strong>to</strong> female gametangia (normally about<br />

1:1) was very high, with less than one female<br />

per 1000 male gametangia. It was concluded<br />

from these experiments that, since intermediate<br />

gametangial arrangements are found in hybrid<br />

haploids, this arrangement is not under the<br />

control of a single pair of non-duplicated allelic<br />

genes, but that a fairly large number of genes<br />

must be involved. Hybridization in some way<br />

upsets the mechanism which controls the<br />

arrangement of gametangia in the parental<br />

species. By treating meiospores of A. macrogynus<br />

with DNA extracted from game<strong>to</strong>thallic cultures<br />

of A. arbuscula, Ojha and Turian (1971) have<br />

demonstrated an inversion of the normal gametangial<br />

arrangement, i.e. a proportion of the<br />

DNA-treated meiospores developed colonies<br />

with hypogynous antheridia instead of the<br />

normal epigynous arrangement. Similar inversions<br />

were also obtained in converse experiments.<br />

In an isolate of the naturally occurring<br />

hybrid A. javanicus, Ji and Dayal (1971) have<br />

shown that although copulation between anisogamous<br />

gametes results in the formation of<br />

sporothalli bearing thin-walled and thick-walled<br />

sporangia, the swarmers from the thick-walled<br />

sporangia rarely develop in<strong>to</strong> game<strong>to</strong>thalli, but<br />

in<strong>to</strong> sporothalli. This is not surprising for a<br />

hybrid, and is possibly due <strong>to</strong> a failure of meiosis<br />

in the thick-walled sporangia.<br />

Sub-genus Cys<strong>to</strong>genes<br />

A life cycle different from Eu-Allomyces is found<br />

in Allomyces moniliformis and A. neo-moniliformis.<br />

There is no independent game<strong>to</strong>thallic generation,<br />

but this stage is probably represented by<br />

a cyst (C. M. Wilson, 1952). The asexual thalli<br />

resemble those of subgenus Eu-Allomyces, bearing


160 CHYTRIDIOMYCOTA<br />

both thin-walled mi<strong>to</strong>sporangia and brown,<br />

thick-walled, pitted meiosporangia. The mi<strong>to</strong>spores<br />

encyst and germinate <strong>to</strong> form a further<br />

crop of asexual thalli. In the meiosporangium,<br />

meiosis takes place, but before cy<strong>to</strong>plasmic<br />

cleavage occurs, the haploid nuclei pair, the<br />

paired nuclei being united by a common nuclear<br />

cap. When cleavage does occur it therefore<br />

results in the formation of some 30 binucleate<br />

cells. When the meiosporangial wall cracks open,<br />

the binucleate cells are released as amoeboid<br />

bodies which may or may not bear flagella, and<br />

it is these cells which form the cysts. A mi<strong>to</strong>tic<br />

division in each cyst results in four haploid<br />

nuclei, and cy<strong>to</strong>plasmic cleavage gives rise <strong>to</strong><br />

four colourless uniflagellate isogametes. These<br />

copulate <strong>to</strong> form biflagellate zygotes, each of<br />

which can develop in<strong>to</strong> an asexual sporothallus.<br />

In the Cys<strong>to</strong>genes life cycle there is thus a freeliving<br />

diploid asexual sporothallic generation,<br />

whereas the haploid generation is reduced <strong>to</strong> the<br />

cysts and gametes.<br />

Sub-genus Brachy-Allomyces<br />

In certain isolates of Allomyces which have been<br />

placed in a ‘form species’ A. anomalus, there are<br />

neither sexual thalli nor cysts. Asexual thalli<br />

bear mi<strong>to</strong>sporangia and brown resting sporangia.<br />

The spores from the resting sporangia<br />

develop directly <strong>to</strong> give asexual thalli again.<br />

The cy<strong>to</strong>logical explanation proposed by C. M.<br />

Wilson (1952) for this unusual behaviour is that,<br />

due <strong>to</strong> complete or partial failure of chromosome<br />

pairing in the resting sporangia, meiosis<br />

does not occur and nuclear divisions are mi<strong>to</strong>tic.<br />

Consequently the zoospores produced from resting<br />

sporangia are diploid, like their parent thalli<br />

and, on germination, give rise <strong>to</strong> diploid asexual<br />

thalli again. Similar failures in chromosome<br />

pairing were also encountered in the hybrids<br />

between A. arbuscula and A. macrogynus leading <strong>to</strong><br />

very low meiospore viability from certain crosses.<br />

In view of this it seemed possible that some of<br />

the forms of A. anomalus might have arisen<br />

through natural hybridization. In a later study,<br />

Wilson and Flanagan (1968) showed that there<br />

is a second way in which the life cycle of<br />

this fungus is maintained without a sexual<br />

phase. In certain isolates, meiosis does occur in<br />

the resting sporangia, followed by apomixis,<br />

i.e. the fusion of two meiosis-derived nuclei in<br />

the same thallus. Propagules from the resting<br />

sporangia are therefore diploid and the cysts<br />

develop in<strong>to</strong> sporothalli. By germinating resting<br />

sporangia in dilute K 2 HPO 4 , a small percentage<br />

of zoospores were produced which developed<br />

in<strong>to</strong> game<strong>to</strong>thalli, some of which were identified<br />

as A. macrogynus and some as A. arbuscula. No<br />

hybrids were found. Thus A. anomalus is not<br />

a single species, but represents sporothalli of<br />

these two species in which the normal alternation<br />

of generations has been upset by cy<strong>to</strong>logical<br />

deviations.<br />

6.5.3 Blas<strong>to</strong>cladiella<br />

About a dozen species of Blas<strong>to</strong>cladiella have been<br />

isolated from soil or water, and one is parasitic<br />

on the cyanobacterium Anabaena (Canter &<br />

Willoughby, 1964). The form of the thallus is<br />

comparatively simple, resembling that of some<br />

monocentric chytrids. There is an extensive<br />

branched rhizoidal system which is attached<br />

either <strong>to</strong> a sac-like sporangium or <strong>to</strong> a cylindrical<br />

trunk-like region bearing a single sporangium<br />

at the tip. In B. emersonii it has been shown that<br />

the rhizoids are chemotropic and function<br />

not only in attachment, but in absorption and<br />

selective translocation of nutrients (Kropf &<br />

Harold, 1982).<br />

Different species of Blas<strong>to</strong>cladiella have life<br />

cycles resembling those of the three subgenera<br />

of Allomyces, and Karling (1973) has proposed<br />

that Blas<strong>to</strong>cladiella should similarly be divided<br />

in<strong>to</strong> three subgenera, i.e. Eucladiella corresponding<br />

<strong>to</strong> Eu-Allomyces, Cys<strong>to</strong>cladiella corresponding<br />

<strong>to</strong> Cys<strong>to</strong>genes, and Blas<strong>to</strong>cladiella corresponding<br />

<strong>to</strong> Brachy-Allomyces. In some species there is an<br />

isomorphic alternation of generations, probably<br />

matching in essential features the Eu-Allomyces<br />

pattern, but cy<strong>to</strong>logical details are needed <strong>to</strong><br />

confirm this. For example, in Blas<strong>to</strong>cladiella<br />

variabilis two kinds of asexual thallus are found.<br />

One bears thin-walled zoosporangia which<br />

release posteriorly uniflagellate swarmers.<br />

These swarmers may develop <strong>to</strong> form thalli<br />

resembling their parents or may give rise <strong>to</strong><br />

the second type of asexual thallus bearing


BLASTOCLADIALES<br />

161<br />

a thick-walled dark-brown sculptured resting<br />

sporangium within the terminal sac. The resting<br />

sporangium releases posteriorly uniflagellate<br />

swarmers which, after swimming, germinate<br />

<strong>to</strong> form sexual thalli of two kinds. About half<br />

of the sexual thalli are colourless (‘female’),<br />

and about half are orange-coloured (‘male’).<br />

However, in contrast <strong>to</strong> the anisogamy of<br />

Eu-Allomyces, inBlas<strong>to</strong>cladiella there is no distinction<br />

in size between the gametes. The orange<br />

and colourless gametes pair <strong>to</strong> produce zygotes,<br />

which germinate directly <strong>to</strong> produce asexual<br />

thalli. In other species (e.g. B. cys<strong>to</strong>gena) the life<br />

cycle is of the Cys<strong>to</strong>genes type, i.e. there are no<br />

game<strong>to</strong>thalli.<br />

In yet other species there is no clear evidence<br />

of sexual fusion. In B. emersonii (Fig. 6.23), the<br />

resting sporangial thallus contains a single<br />

globose, dark reddish brown resting sporangium<br />

with a dimpled wall. Meiosis occurs during<br />

development of the resting sporangium (Olson<br />

& Reichle, 1978). After a resting period, the wall<br />

cracks open and one <strong>to</strong> four papillae protrude<br />

from which swarmers are released. The swarmers<br />

germinate <strong>to</strong> form two types of thallus bearing<br />

thin-walled zoosporangia. About 98% of the<br />

swarmers give rise <strong>to</strong> thalli bearing colourless<br />

sporangia (Fig. 6.23a), and about 2% <strong>to</strong> thalli<br />

with sporangia coloured orange due <strong>to</strong> the<br />

presence of g-carotene. The colourless thalli<br />

develop rapidly and are ready <strong>to</strong> discharge<br />

zoospores within 24 h. These have about twice<br />

the DNA content as the swarmers released from<br />

resting sporangia (Horgen et al., 1985). Thus<br />

young colourless thalli are at first haploid, but<br />

release diploid zoospores. The manner in which<br />

the diploid state of the colourless thalli or of the<br />

resting sporangia is brought about is not known.<br />

The life cycle of B. emersonii thus corresponds<br />

<strong>to</strong> that of the sub-genus Brachyallomyces.<br />

Blas<strong>to</strong>cladiella emersonii has a number of other<br />

unusual features. If zoospore suspensions are<br />

pipetted on<strong>to</strong> yeast pep<strong>to</strong>ne glucose (YPG)<br />

agar, the majority of thalli which develop will<br />

be of the thin-walled colourless type. On the<br />

same medium containing 10 mM bicarbonate,<br />

resting sporangial thalli develop. The addition<br />

of 40 80 mM KCl, NaCl or NH 4 Cl, or exposure<br />

of cultures <strong>to</strong> ultra-violet light, will similarly<br />

induce the formation of resting sporangia<br />

(Horgen & Griffin, 1969). Thus, by means of<br />

simple manipulation of the environment it is<br />

possible <strong>to</strong> switch the metabolic activities of<br />

the fungus in<strong>to</strong> one of two morphogenetic<br />

Fig 6.23 Blas<strong>to</strong>cladiella emersonii. (a) Thin-walled thallus releasing zoospores. (b) Three-day-old thallus with immature resting<br />

sporangium. (c) Thallus with germinating resting sporangium showing the cracked wall and four exit tubes. (d) Zoospores from<br />

thin-walled thallus.


162 CHYTRIDIOMYCOTA<br />

pathways. There are important differences in the<br />

activities of certain enzymes (Cantino et al., 1968;<br />

Lovett, 1975). In the absence of bicarbonate,<br />

there is evidence for the operation of a tricarboxylic<br />

acid cycle, whereas in the presence of<br />

bicarbonate, part of this cycle is reversed, leading<br />

<strong>to</strong> alternative pathways of primary carbon<br />

metabolism. In addition, a polyphenol oxidase,<br />

absent in the thin-walled thallus, replaces the<br />

normal cy<strong>to</strong>chrome oxidase. There is also<br />

increased synthesis of melanin and of chitin in<br />

the presence of bicarbonate. The effect of<br />

bicarbonate can be brought about by increased<br />

levels of CO 2 .<br />

Another unusual feature is that B. emersonii<br />

fixes CO 2 more rapidly in the light than in the<br />

dark. In the presence of CO 2 , light-grown thalli<br />

show a number of differences when compared<br />

with dark-grown controls. Illuminated thalli take<br />

about three hours longer <strong>to</strong> mature, and are<br />

larger than dark-grown thalli. They also have an<br />

increased rate of nuclear division and a higher<br />

nucleic acid content. The most effective wavelengths<br />

for this increased CO 2 fixation (or<br />

lumisynthesis) lie between 400 and 500 nm, i.e.<br />

at the blue end of the spectrum. This suggests<br />

that the pho<strong>to</strong>recep<strong>to</strong>r should be a yellowish<br />

substance. Attempts <strong>to</strong> identify the pho<strong>to</strong>recep<strong>to</strong>r<br />

have as yet been unsuccessful, but it is known<br />

not <strong>to</strong> be a carotenoid.<br />

6.6 Monoblepharidales<br />

This group includes about 20 species and is<br />

represented by 5 genera, namely Monoblepharis,<br />

Monoblepharella, Gonapodya, Oedogoniomyces and<br />

Harpochytrium. <strong>Fungi</strong> belonging <strong>to</strong> this order<br />

can be isolated from soil samples or from twigs<br />

or fruits submerged in freshwater, sometimes<br />

under anoxic conditions (Karling, 1977; Fuller<br />

& Clay, 1993). Whisler (1987) has given details<br />

of isolation techniques. Most species are saprotrophs<br />

and several are available in culture. In all<br />

genera the thallus is eucarpic either with<br />

rhizoids or a holdfast, and with branched or<br />

unbranched filaments. The walls contain<br />

microfibrils of chitin (Bartnicki-Garcia, 1968),<br />

but the walls of G. prolifera also contain cellulose<br />

(Fuller & Clay, 1993). A characteristic feature is<br />

the frothy or alveolate appearance of the<br />

cy<strong>to</strong>plasm caused by the presence of numerous<br />

vacuoles often arranged in a regular pattern.<br />

Asexual reproduction is by posteriorly uniflagellate<br />

zoospores which are borne in terminal,<br />

cylindrical or flask-shaped sporangia. Sexual<br />

reproduction, where known, is unique for fungi<br />

in being oogamous with a large egg and a<br />

smaller, posteriorly flagellate sperma<strong>to</strong>zoid. The<br />

egg may be retained within the oogonium or<br />

may move <strong>to</strong> its mouth by amoeboid movement<br />

in some species of Monoblepharis, or propelled by<br />

the lashing of the flagellum of the sperma<strong>to</strong>zoid<br />

in Monoblepharella and Gonapodya.<br />

6.6.1 The zoospore<br />

The fine structure of zoospores is similar in<br />

representatives of all five genera (Fig. 6.24; see<br />

Mollicone & Longcore, 1994, 1999). In all cases<br />

the body of the zoospore is oval, the narrow<br />

part facing forward and with a long whiplash<br />

flagellum trailing from the wider posterior.<br />

Amoeboid changes of shape may occur and<br />

swimming zoospores may develop pseudopodia<br />

anteriorly. The body of the zoospore is differentiated<br />

in<strong>to</strong> three regions: an anterior region<br />

which is often devoid of organelles apart from<br />

lipid globules, a few vacuoles and tubular<br />

cisternae; a central region which contains the<br />

nucleus, surrounded by ribosomal aggregations<br />

(sometimes termed the nuclear cap), microbodies<br />

and spherical mi<strong>to</strong>chondria with flattened<br />

cristae; and a posterior ‘foamy’ region at<br />

the base of which are the functional kine<strong>to</strong>some,<br />

a non-functional kine<strong>to</strong>some and a rumposomal<br />

complex. The functional kine<strong>to</strong>some is surrounded<br />

by a striated disc, apparently anchored<br />

<strong>to</strong> annular cisternae. From an electron-dense<br />

region of the striated disc, about 31 34 microtubules<br />

extend outwards in<strong>to</strong> the body of the<br />

zoospore. Water expulsion vacuoles have been<br />

identified in the anterior part of the zoospore<br />

of G. prolifera. Another distinctive feature in this<br />

fungus is the presence of a pair of paraxonemal<br />

structures, solid cylindrical fibres which are


MONOBLEPHARIDALES<br />

163<br />

Fig 6.24 Summary diagram of zoospore ultrastructure<br />

of Monoblepharis polymorpha. Abbreviations: angular cisternae<br />

(ac), endoplasmic reticulum (er), kine<strong>to</strong>some (K), lipid globule<br />

(L), mi<strong>to</strong>chondria (M), microbody (mb), microtubule (mt),<br />

nucleus (N), non-flagellated centriole (nfc), transition zone<br />

plug (O), kine<strong>to</strong>some prop (P), ribosomal aggregate (R),<br />

rumposome (Ru), striated disc (sd), vacuole (V).The diameter<br />

of the zoospore is about 6 mm. Reprinted with permission<br />

from Mollicone and Longcore (1994), Mycologia.<br />

ß The Mycological Society of America.<br />

smaller in diameter than the axonemal microtubules,<br />

running parallel <strong>to</strong> them within the<br />

axoneme and connected at intervals <strong>to</strong> doublets<br />

3 and 8 (Mollicone & Longcore, 1999).<br />

6.6.2 Monoblepharis<br />

Species of Monoblepharis occur in quiet silt-free<br />

pools containing neutral or slightly alkaline<br />

water (i.e. pH 6.4 7.5) on waterlogged twigs on<br />

which the bark is still present. Twigs of birch,<br />

ash, elm and especially oak are suitable substrata,<br />

and although samples taken at varying<br />

times throughout the year may yield growths<br />

of the fungus, there are two main periods of<br />

vegetative growth, one in spring and another<br />

in autumn, with resting periods during the<br />

summer and winter months. Low temperatures<br />

appear <strong>to</strong> favour asexual development and good<br />

growth can be obtained on twigs incubated in<br />

dishes of distilled water at temperatures around<br />

3°C. The mycelium is delicate and vacuolate.<br />

The hyphae are multinucleate. During the<br />

formation of a sporangium, a multinucleate<br />

tip is cut off by a septum. The cy<strong>to</strong>plasm cleaves<br />

around the nuclei <strong>to</strong> form zoospore initials<br />

which are at first angular and then later pearshaped.<br />

The ripe sporangium is cylindrical or<br />

club-shaped and may not be much wider than<br />

the hypha bearing it. A pore is formed at<br />

the tip of the sporangium through which the<br />

zoospores escape by amoeboid crawling. The<br />

free zoospores swim away. On coming <strong>to</strong> rest,<br />

a zoospore encysts and germinates by emitting<br />

a germ tube. The single nucleus of the zoospore<br />

cyst divides and further nuclear divisions occur<br />

as the germ tube elongates.<br />

Sexual reproduction can be induced by<br />

incubating twigs at room temperature. Light<br />

also affects reproduction in M. macrandra.<br />

Cultures of this fungus incubated in the dark<br />

produced only game<strong>to</strong>thalli whilst those grown<br />

in light formed only sporothalli (Marek, 1984).<br />

In M. polymorpha and related species, the antheridia<br />

are epigynous, becoming cut off by a basal<br />

septum. Beneath the antheridium the hypha<br />

becomes swollen somewhat asymmetrically so<br />

that the antheridium is displaced in<strong>to</strong> a lateral<br />

position. The swollen subterminal part becomes<br />

spherical and is then cut off by a basal septum<br />

<strong>to</strong> form the oogonium. In M. sphaerica and<br />

some other species, the arrangement of the sex<br />

organs is the reverse of that in M. polymorpha,<br />

i.e. hypogynous. In M. macrandra the antheridia<br />

and oogonia may grow as solitary organs at the


164 CHYTRIDIOMYCOTA<br />

Fig 6.25 Monoblepharis macrandra reproduction. (a) Terminal<br />

zoosporangium containing cleaved zoospores. (b) Solitary<br />

terminal antheridium. (c) Solitary terminal oogonium with<br />

apical receptive area. (d) Oogonium with hypogynous<br />

antheridium. (e) Sperma<strong>to</strong>zoid release from solitary terminal<br />

antheridium. (f) Exogenous oospore on empty oogonium with<br />

bullations on the wall of the oogonium and lipid inclusions<br />

in the cy<strong>to</strong>plasm. (a e) <strong>to</strong> same scale.Traced from Whisler<br />

and Marek (1987), with permission by Southeastern Publishing<br />

Corporation.<br />

tips of the hyphae (Figs. 6.25b,c) or in pairs,<br />

with the antheridium in a hypogynous position<br />

(Fig. 6.25d). The antheridium often releases<br />

sperm before the adjacent oogonium is ripe.<br />

Each antheridium forms about four <strong>to</strong> eight<br />

posteriorly uniflagellate swarmers which resemble,<br />

but are somewhat smaller than, the zoospores.<br />

The oogonium contains a single spherical<br />

uninucleate oosphere, and when this is mature<br />

an apical receptive papilla on the oogonial wall<br />

breaks down. A sperma<strong>to</strong>zoid approaching the<br />

receptive papilla of the oogonium becomes<br />

caught up in mucus and fusion with the<br />

oosphere then follows, the flagellum of the<br />

sperma<strong>to</strong>zoid being absorbed within a few<br />

minutes. Following plasmogamy, the oospore<br />

secretes a golden-brown wall around itself<br />

and nuclear fusion later occurs. In some species,<br />

e.g. M. sphaerica, the oospore remains within<br />

the oogonium (endogenous) but in others, e.g.<br />

M. macrandra and M. polymorpha, the oospore<br />

begins <strong>to</strong> move <strong>to</strong>wards the mouth of the<br />

oogonium within a few minutes of fertilization,<br />

and remains exogenous, i.e. attached <strong>to</strong> it<br />

(Fig. 6.25f). In the exogenous species, nuclear<br />

fusion is delayed but finally fusion occurs and<br />

the oospore becomes uninucleate. In some<br />

species the oospore wall remains smooth, but<br />

in others such as M. macrandra the wall may be<br />

ornamented by hemispherical warts or bullations<br />

(Fig. 6.25f). The oospore germinates after<br />

a resting period which coincides with frozen<br />

winter conditions or summer drought by producing<br />

a single hypha which branches <strong>to</strong> form<br />

a mycelium. The cy<strong>to</strong>logical details of the life<br />

cycle are not fully known but it seems likely<br />

that reduction division occurs during the germination<br />

of the overwintered oospores.


7<br />

Zygomycota<br />

7.1 <strong>Introduction</strong><br />

The phylum Zygomycota comprises the first<br />

group of fungi considered in this book which<br />

lacks any motile stage. Asexual reproduction is<br />

by spores which are called aplanospores because<br />

they are non-motile, and sporangiospores<br />

because they are typically contained within<br />

sporangia. They are dispersed passively by wind,<br />

insects and rain splash, although violent liberation<br />

of entire sporangia (e.g. Pilobolus) or individual<br />

spores (e.g. Basidiobolus, En<strong>to</strong>mophthora) can<br />

also occur. Sexual reproduction is by gametangial<br />

copulation which is typically isogamous and<br />

results in the formation of a zygospore. The<br />

mycelial organization is coenocytic, and the<br />

cell wall contains chitin and its deacetylated<br />

derivative, chi<strong>to</strong>san (Bartnicki-Garcia, 1968, 1987;<br />

see Fig. 1.5). As in the Chytridio-, Asco- and<br />

Basidiomycota, the mi<strong>to</strong>chondria possess lamellate<br />

cristae, and the Golgi system is reduced <strong>to</strong><br />

single cisternae. Lysine is synthesized by the<br />

a-aminoadipic acid (AAA) route, as it appears <strong>to</strong><br />

be in all Eumycota.<br />

General accounts of the Zygomycota have<br />

been given by Benjamin (1979), Benny (2001) and<br />

Benny et al. (2001). Molecular evidence indicates<br />

that the group may have diverged from the<br />

Chytridiomycota early in the his<strong>to</strong>ry of terrestrial<br />

life. The Zygomycota, in turn, probably<br />

gave rise <strong>to</strong> the Asco- and Basidiomycota, i.e.<br />

the ‘higher fungi’ (Jensen et al., 1998; Schüssler<br />

et al., 2001). Two classes are included in the<br />

Zygomycota, namely Zygomycetes comprising<br />

870 species in 10 orders, and Trichomycetes<br />

with 218 species in 3 orders (Kirk et al., 2001).<br />

The most prominent orders of the Zygomycetes<br />

are the Mucorales, En<strong>to</strong>mophthorales and<br />

Glomales. Mucorales are ubiqui<strong>to</strong>us in soil and<br />

dung mostly as saprotrophs, although a few are<br />

parasitic on plants and animals. En<strong>to</strong>mophthorales<br />

include a number of insect parasites, but<br />

some saprotrophic forms also exist. Glomales<br />

are mutualistic symbionts associated with<br />

almost all kinds of terrestrial plants as arbuscular<br />

and vesicular arbuscular mycorrhiza. Trichomycetes<br />

are mostly commensal in the guts of<br />

arthropods, e.g. millipedes and the larvae of<br />

aquatic insects.<br />

The Zygomycetes are almost certainly polyphyletic,<br />

but the precise evolutionary relationships<br />

within this class are still controversial<br />

(O’Donnell et al., 2001; Schüssler et al., 2001;<br />

Tanabe et al., 2004, 2005), and comparisons of<br />

numerous representative organisms with several<br />

different DNA sequences, e.g. genes encoding<br />

ribosomal RNA, cy<strong>to</strong>chrome oxidase or cy<strong>to</strong>skeletal<br />

proteins, will be required before a satisfac<strong>to</strong>ry<br />

natural arrangement can be found. A recent<br />

phylogenetic scheme is presented in Fig. 7.1.<br />

7.2 Zygomycetes: Mucorales<br />

In most members of the Mucorales, numerous<br />

spores are contained in globose sporangia borne<br />

at the tips of aerial sporangiophores (Fig. 7.2).<br />

Within the sporangium the spores may surround


166 ZYGOMYCOTA<br />

Fig 7.1 Recent phylogenetic scheme of<br />

the Zygomycota based on partial<br />

sequences of the gene encoding a subunit<br />

of RNA polymerase II.Orders discussed in<br />

detail in this book include Mucorales<br />

(Sections 7.2 7.3), Zo opagales<br />

(Section 7.4), En<strong>to</strong>mophthorales<br />

(Section 7.5) in the Zygomycetes, and<br />

Harpellales (Section 7.7) in the<br />

Trichomycetes.The Glomales (Section 7.6),<br />

only distantly related <strong>to</strong> other<br />

Zygomycota, were not included in this<br />

analysis. Redrawn and modified from<br />

Tanabe et al. (2004), with permission<br />

from Elsevier.<br />

a central core or columella, although in some<br />

species (e.g. Mortierella spp.) the columella is<br />

greatly reduced. Some species possess fewspored<br />

sporangia, termed sporangiola, which<br />

are often dispersed as a unit, and in some groups<br />

the spores are arranged as a single row inside a<br />

cylindrical sac termed a merosporangium. Yet<br />

other Mucorales reproduce by means of unicellular<br />

propagules which are sometimes termed<br />

conidia, but Benjamin (1979), in his consideration<br />

of asexual propagules formed in the<br />

Zygomycota, has recommended the use of the<br />

term ‘sporangiolum’ instead of ‘conidium’ in<br />

this context. It is believed that ‘conidia’ may have<br />

evolved several times within different groups of<br />

Mucorales from forms with monosporous sporangiola.<br />

A distinction between sporangiospores<br />

and conidia is that germinating sporangiospores<br />

lay down a new wall, continuous with the germ<br />

tube, within their original spore wall, whilst<br />

within germinating conidia there is no new wall<br />

layer formed.<br />

The Mucorales are mostly saprotrophic and<br />

are abundant in soil, on dung and on other<br />

organic matter in contact with the soil. They may<br />

play an important role in the early colonization<br />

of substrata. Sometimes, however, they can<br />

behave as weak pathogens of soft plant tissues,<br />

e.g. Rhizopus s<strong>to</strong>lonifer can cause a rot of sweet<br />

pota<strong>to</strong>es or fruits such as apples, <strong>to</strong>ma<strong>to</strong>es and<br />

strawberries (Plate 3d). Such infections may<br />

cause spoilage of food (Samson et al., 2002).<br />

Some species are parasitic on other fungi, a<br />

common example being Spinellus fusiger which<br />

forms a tuft of sporangiophores on the caps of<br />

moribund fruit bodies of Mycena spp. (Plate 3e).<br />

Others cause diseases of animals including<br />

man, especially patients suffering from diabetes,<br />

leukaemia and cancer. Lesions may be localized<br />

in the brain, lungs or other organs, or may be<br />

disseminated, e.g. at various points in the<br />

vascular system (Kwon-Chung & Bennett, 1992).<br />

Species of Rhizopus and Mucor are reported from<br />

human lesions, and these genera <strong>to</strong>gether with


ZYGOMYCETES: MUCORALES<br />

167<br />

species of Absidia may also infect domestic<br />

animals.<br />

A number of species have been used in the<br />

production of oriental foods such as sufu,<br />

tempeh and ragi (Nout & Aidoo, 2002) and<br />

some are used as starters in the saccharification<br />

of starchy materials before fermentation <strong>to</strong><br />

alcohol (Hesseltine, 1991). In modern biotechnology,<br />

many mucoralean fungi are employed in<br />

biotransformation processes (for references, see<br />

Kieslich, 1997). Further, a number of species are<br />

oleaginous, i.e. they are able <strong>to</strong> synthesize and<br />

accumulate lipids <strong>to</strong> over 20% (dry weight) of<br />

their biomass. Because these lipids (principally<br />

triacylglycerides) may be enriched in polyunsaturated<br />

fatty acids (PUFAs), oleaginous members<br />

of the Mucorales are of current biotechnological<br />

interest (Certik & Shimizu, 1999).<br />

Extensive studies of nutrition and physiology<br />

have been made. A wide variety of sugars can be<br />

used, and whilst starch can be decomposed by<br />

some species, cellulose is generally not utilized.<br />

Under anaerobic conditions, ethanol and numerous<br />

organic acids are produced. Many Mucorales<br />

need an external supply of vitamins for growth<br />

in synthetic culture. Thiamine is a common<br />

requirement, and the amount of growth of<br />

Phycomyces has been used as an assay for the<br />

concentration of thiamine.<br />

Zycha et al. (1969) have given a general<br />

account of the taxonomy of the Mucorales,<br />

including keys <strong>to</strong> genera and species. Benny<br />

et al. (2001) recognized 13 families and 57 genera.<br />

Classification and identification are based<br />

largely on the morphology of the anamorph.<br />

However, DNA sequence comparisons indicate<br />

that several families and even some larger<br />

genera are polyphyletic (O’Donnell et al., 2001),<br />

meaning that the traditional family-level classification<br />

scheme is artificial. We retain it here<br />

because it presents an accessible framework<br />

of morphological features within which the<br />

Mucorales can be unders<strong>to</strong>od, and because<br />

convincing alternative schemes have not yet<br />

been put forward.<br />

7.2.1 Growth and asexual reproduction<br />

The mycelium is coarse, coenocytic and<br />

richly branched, the branches tapering <strong>to</strong> fine<br />

Fig 7.2 Mucor mucedo. (a) Mycelium and young<br />

sporangiophores with globules of liquid attached. (b) Immature<br />

sporangium with the columella visible through the sporangial<br />

wall. (c) Dehisced sporangium showing the columella, the frill<br />

representing the remains of the sporangial wall, and<br />

sporangiospores.<br />

points (Fig. 7.2). Later, septa may appear.<br />

Thick-walled mycelial segments (chlamydospores)<br />

may be cut off by such septa (Benjamin,<br />

1979) and in certain species, e.g. Mucor racemosus,<br />

the presence of chlamydospores in sporangiophores<br />

may be a useful diagnostic feature<br />

(Fig. 7.14b). In anaerobic liquid culture, especially<br />

in the presence of CO 2 , several species of<br />

Mucor (e.g. M. rouxii) grow in a yeast-like instead<br />

of a filamen<strong>to</strong>us form (Fig. 7.3) but revert <strong>to</strong><br />

filamen<strong>to</strong>us growth in the renewed presence<br />

of O 2 . The cell walls of Mucorales are chemically<br />

complex (Ruiz-Herrera, 1992; Gooday, 1995).<br />

Chitin microfibrils are present but are often


168 ZYGOMYCOTA<br />

Fig 7.3 Mucor rouxii. (a) Yeast-like growth in liquid medium<br />

under anaerobic conditions 24 h after inoculation with spores.<br />

(b) Filamen<strong>to</strong>us growth from spores in liquid medium under<br />

aerobic conditions 4 h after inoculation.<br />

deacetylated <strong>to</strong> chi<strong>to</strong>san. Other compounds such<br />

as poly-D-glucuronides, polyphosphates, proteins,<br />

lipids, purines, pyrimidines, magnesium and<br />

calcium have also been detected. Comparison of<br />

the structure and composition of yeast-like and<br />

filamen<strong>to</strong>us cells of Mucor rouxii shows that the<br />

yeast-like cells have much thicker walls. They<br />

also have a mannose content about five times<br />

as great as that of filamen<strong>to</strong>us cell walls. The<br />

synthesis of chitin microfibrils takes place<br />

within chi<strong>to</strong>somes which have been described<br />

from sporangiophores of Phycomyces (Herrera-<br />

Estrella et al., 1982; Gamow et al., 1987) and<br />

from a number of other fungi with chitinous<br />

walls (see p. 6).<br />

Asexual reproduction is by aplanospores<br />

(sporangiospores) contained in globose or pearshaped<br />

sporangia, which are borne singly at<br />

the tip of a sporangiophore or on branched<br />

sporangiophores. In Absidia (Fig. 7.17), the<br />

sporangia are arranged in whorls on aerial<br />

branches, and in many species of Rhizopus the<br />

sporangiophores arise in groups from a clump<br />

of rhizoids (Fig. 7.16). Sporangiophores are<br />

often pho<strong>to</strong>tropic, and several studies on the<br />

pho<strong>to</strong>tropism of the strikingly large sporangiophores<br />

of Phycomyces blakesleeanus have been<br />

undertaken (see p. 169) and have been summarized<br />

in elegant and stimulating reviews of the<br />

biology of this fungus (Bergman et al., 1969;<br />

Cerdá-Olmedo & Lipson, 1987; Cerdá-Olmedo,<br />

2001). ‘The sporangiophore of the fungus<br />

Phycomyces is a gigantic, single-celled, erect,<br />

cylindrical aerial hypha. It is sensitive <strong>to</strong> at<br />

least four distinct stimuli: light, gravity,<br />

stretch, and some unknown stimulus by which<br />

it avoids solid objects. These stimuli control a<br />

common output, the growth rate, producing<br />

either temporal changes in the growth rate or<br />

tropic responses’ (Bergman et al., 1969). The<br />

avoidance by the sporangiophore of solid<br />

objects is termed the avoidance response or<br />

fugitropism. Despite its obvious fascination, the<br />

mechanisms behind the avoidance response are<br />

still not unders<strong>to</strong>od. Because, under certain<br />

conditions, P. blakesleeanus may also develop<br />

much smaller sporangiophores (microsporangiophores<br />

or microphores), the larger sporangiophores<br />

are sometimes termed macrophores.<br />

Despite their remarkable height (Fig. 7.4), for<br />

much of their length the macrophores are<br />

a constant 100 mm in diameter. The wall, about<br />

0.6 mm thick, encloses a peripheral layer of<br />

cy<strong>to</strong>plasm of about 30 mm surrounding a<br />

central vacuole about 40 mm in diameter<br />

(Fig. 7.5). The mature sporangium is spherical<br />

and some 500 mm across.<br />

The sporangiophore of Phycomyces develops<br />

as a conical outgrowth from the vegetative


ZYGOMYCETES: MUCORALES<br />

169<br />

mycelium. Elongation of the sporangiophore is<br />

confined <strong>to</strong> a yellow-pigmented growing zone<br />

(about 1 cm long) beneath the apex. In the<br />

absence of light, sporangiophores grow vertically<br />

as a negative response <strong>to</strong> gravity. Schimek et al.<br />

(1999) have suggested that gravity may be<br />

detected by a combination of at least two<br />

mechanisms. Proteinaceous crystals located<br />

inside vacuoles have a higher density than the<br />

vacuolar sap and therefore sediment in response<br />

<strong>to</strong> gravity, whereas a cluster of buoyant lipid<br />

droplets less dense than the cy<strong>to</strong>plasm floats <strong>to</strong><br />

the apex of the sporangiophore. Both mechanisms<br />

would be different from that found in<br />

the fruit bodies of basidiomycetes such as<br />

Flammulina, in which nuclei denser than the<br />

surrounding cy<strong>to</strong>plasm seem <strong>to</strong> be the organelles<br />

involved in graviperception (see p. 546).<br />

7.2.2 Pho<strong>to</strong>tropism in Phycomyces<br />

If a sporangiophore is subjected <strong>to</strong> unilateral<br />

illumination it bends <strong>to</strong>wards the light,<br />

especially blue light. Pho<strong>to</strong>tropism in Phycomyces<br />

is extremely sensitive, the lower threshold being<br />

1nWm 2 , which is equivalent <strong>to</strong> the light<br />

emitted by a single star at night (Cerdá-<br />

Olmedo, 2001). Bending is the consequence of<br />

a deceleration of about 6% in the growth rate<br />

of the side proximal <strong>to</strong> the direction of light,<br />

and an increase by the same rate on the distal<br />

side (Fig. 7.4). Because the refractive index of<br />

the sporangiophore contents exceeds that of air,<br />

the sporangiophore functions as a cylindrical<br />

lens, focusing unilateral light on the distal<br />

wall of the sporangiophore, resulting in more<br />

intense illumination of that side. Evidence in<br />

support of the lens effect is the demonstration<br />

that sporangiophores immersed in mineral oil<br />

with a higher refractive index than that of the<br />

sporangiophore contents function as a diverging<br />

lens and bend away from the light. The illumination<br />

of the edge of a sporangiophore by a narrow<br />

beam of light from a laser is followed by bending<br />

of the sporangiophore in a direction perpendicular<br />

<strong>to</strong> the light beam (Meistrich et al., 1970).<br />

Pho<strong>to</strong>recep<strong>to</strong>rs are located in the plasma<br />

membrane (Fukshansky, 1993), and the transmission<br />

of the signal leads <strong>to</strong> localized wall<br />

softening and the synthesis of new cell wall<br />

Fig 7.4 Sporangiophore development of Phycomyces<br />

blakesleeanus in standard test tubes (about1.5 cm diameter).<br />

The tubes were wrapped except for the tip of tube (a) or a<br />

square on the right-hand side near the <strong>to</strong>p of tube (b) In tube<br />

(a), the sporangiophores have grown straight <strong>to</strong>wards the<br />

light, whereas in tube (b) they have bent <strong>to</strong>wards the lateral<br />

light source.<br />

material (Herrera-Estrella & Ruiz-Herrera, 1983;<br />

Ortega, 1990).<br />

A central problem in studies of pho<strong>to</strong>responses<br />

is the nature of the pho<strong>to</strong>recep<strong>to</strong>r(s).<br />

Two pho<strong>to</strong>recep<strong>to</strong>rs one for low and the<br />

other for high light intensities are involved<br />

in determining the pho<strong>to</strong>tropism in Phycomyces,<br />

and there are also two recep<strong>to</strong>rs each for<br />

light-induced microphore formation, macrophore<br />

formation, and carotenoid biosynthesis<br />

(Cerdá-Olmedo, 2001). A clue <strong>to</strong> the possible


170 ZYGOMYCOTA<br />

Fig 7.5 Sporangiophore of<br />

Phycomyces as a cylindrical lens.<br />

Upper portion: light ray L impinges<br />

from the left and is refracted at<br />

the first surface.The ratio of<br />

‘path length’of the light ray in<br />

the proximal part of the<br />

sporangiophore <strong>to</strong> the path length<br />

in the distal part is PM/DM.The<br />

maximum value of the angle b is<br />

about 20°.Lowerportion:<br />

sporangiophore in section <strong>to</strong> show<br />

the peripheral layer of cy<strong>to</strong>plasm<br />

surrounding the central vacuole.<br />

The values are estimates of the<br />

refractive index of cy<strong>to</strong>plasm and<br />

vacuolar sap. Diagram modified<br />

from Bergman et al. (1969).<br />

nature of the pho<strong>to</strong>recep<strong>to</strong>rs can be obtained by<br />

studying the action spectrum of the response<br />

over a range of light wavelengths. The pho<strong>to</strong>tropic<br />

curvature of Phycomyces sporangiophores<br />

has a similar action spectrum <strong>to</strong> the growth<br />

response of the vegetative mycelium, which is<br />

also stimulated by light. There are several clearly<br />

defined peaks at 485, 455, 385 and 280 nm, i.e.<br />

mostly in the blue part of the spectrum.<br />

Although b-carotene is present in large amounts<br />

in the growing zone of the sporangiophore, the<br />

pho<strong>to</strong>recep<strong>to</strong>r system is more likely <strong>to</strong> comprise<br />

a flavin-type molecule and a pterin-type protein<br />

(Flores et al., 1999; Galland & Tölle, 2003).<br />

Mutants with less than 0.1% of the wild-type b-<br />

carotene content remain fully pho<strong>to</strong>sensitive.<br />

However, b-carotene is involved in the other<br />

light-induced responses. The signalling chains<br />

involved in transduction of the light signal are<br />

only partially unravelled at present (Cerdá-<br />

Olmedo, 2001).<br />

Moss and Baker (2002) have described a<br />

technique for demonstrating the pho<strong>to</strong>tropic<br />

response of P. blakesleeanus in the labora<strong>to</strong>ry.<br />

7.2.3 Sporangiophore development in<br />

Phycomyces<br />

As the sporangiophore of Phycomyces develops it<br />

rotates. Castle (1942) followed the growth and<br />

rotation of the sporangiophore by attaching<br />

Lycopodium spores as markers and tracking the<br />

displacement of the markers. His findings are<br />

illustrated in Fig. 7.6. After a period of apical<br />

growth of the tubular sporangiophore (stage I),<br />

the sporangium appears as a terminal swelling<br />

and growth ceases. During this period (stage II),<br />

growth is limited <strong>to</strong> sporangial enlargement.<br />

In the next period (stage III) no further enlargement<br />

of the sporangium occurs, and elongation<br />

is also at a standstill. During stages IVA and IVB,<br />

elongation of the sporangiophore is resumed and<br />

growth is mainly localized in a zone somewhat<br />

below the sporangium. During stage I the tip of<br />

the sporangiophore rotates clockwise (as seen<br />

from above looking down) through a maximum<br />

angle of about 90°. There is no rotary movement<br />

during stages II and III. When sporangiophore<br />

elongation recommences in stage IVA, the direction<br />

of rotation is now anti-clockwise (as seen from


ZYGOMYCETES: MUCORALES<br />

171<br />

Fig 7.6 Diagram of developmental stages of<br />

the sporangiophore of Phycomyces. Regions<br />

in which growth is taking place are stippled.<br />

The rotary component of growth is indicated.<br />

During stage I the axis of growth is directed<br />

sinistrally, in stages II and III growth is<br />

unoriented. In Stage IVA dextral spiralling<br />

occurs and in Stage IVB sinistral spiralling again<br />

takes place.<br />

above). During this stage, which lasts about an<br />

hour, markers attached <strong>to</strong> the growth zone may<br />

make up <strong>to</strong> two complete revolutions around the<br />

axis. During stage IVB the direction of rotation<br />

reverses once more. The reasons for the spiral<br />

growth are far from clear (see Ortega et al., 2003).<br />

It is known that the chitin microfibrils which<br />

make up the wall of the sporangiophore show<br />

a right-handed or Z-spiral orientation. One<br />

possible explanation is that the laying down of<br />

the fibrils in this way is responsible for the<br />

rotation. A second is that the extension due <strong>to</strong><br />

turgor pressure of a cylinder whose walls are<br />

composed of spirally arranged fibrils would<br />

naturally result in a passive rotation. The<br />

phenomenon of spiral growth is not peculiar <strong>to</strong><br />

Phycomyces, occurring also during elongation of<br />

the sporangiophores of other members of the<br />

Mucorales such as Thamnidium and Pilobolus, and<br />

in various cylindrical plant cells.<br />

The mechanical properties of the sporangiophore<br />

of Phycomyces change during development.<br />

During stage II, when no elongation of the<br />

sporangiophore is taking place, the sporangiophore<br />

shows elastic deformation when small<br />

loads are applied <strong>to</strong> it, i.e. the fractional change<br />

in length is directly proportional <strong>to</strong> the applied<br />

load, and on removal of the load, the<br />

sporangiophore returns <strong>to</strong> its original length.<br />

During stage IV, although the sporangiophore<br />

changes in length in response <strong>to</strong> applied loads,<br />

upon unloading the sporangiophore does not<br />

return <strong>to</strong> its original length.<br />

There is evidence that the spores secrete<br />

one or more unknown substances which control<br />

elongation of the sporangiophore. If mature<br />

sporangia are removed, growth of the sporangiophore<br />

ceases. Replacement of the detached<br />

sporangium with a substitute sporangium,<br />

with a suspension of spores, or with a drop<br />

of supernatant liquid from a centrifuged spore<br />

suspension, results in resumption of growth.<br />

Another effect of the removal of a ripe sporangium<br />

is that branching is induced in the<br />

sporangiophore, and this phenomenon has been<br />

likened <strong>to</strong> the breaking of apical dominance<br />

upon removal of the terminal bud in shoots of<br />

angiosperms.<br />

7.2.4 Sporangium development<br />

The tip of the sporangiophore expands <strong>to</strong> form<br />

the sporangium initial containing numerous<br />

nuclei which continue <strong>to</strong> divide. A dome-shaped<br />

septum is laid down and cuts off a distal portion<br />

which will contain the spores, from a cylindrical<br />

or subglobose spore-free core, the columella.


172 ZYGOMYCOTA<br />

The columella is curved from its inception.<br />

Cleavage planes separate the nuclei within the<br />

sporangium, and finally the spores are cleaved<br />

out. They may be uninucleate or multinucleate<br />

according <strong>to</strong> the species, e.g. M. hiemalis and<br />

Absidia glauca have predominantly uninucleate<br />

spores (S<strong>to</strong>rck & Morrill, 1977) whilst M. mucedo,<br />

P. blakesleeanus, Rhizopus s<strong>to</strong>lonifer and Syzygites<br />

megalocarpus have multinucleate spores<br />

(Hammill & Secor, 1983). The number of spores<br />

formed is very variable. On nutrient-poor media<br />

minute sporangia containing very few spores<br />

may be formed, but in P. blakesleeanus the number<br />

of spores may be as high as 50 000 100 000 in a<br />

single sporangium of normal size.<br />

The ultrastructure of developing sporangia<br />

of Gilbertella has been studied by Bracker<br />

(1968a) and is essentially similar in multisporous<br />

columellate sporangia of other Mucorales,<br />

e.g. M. mucedo (Hammill, 1981) and Zygorhynchus<br />

heterogamus (Edelman & Klomparens, 1994). One<br />

difference is that in M. mucedo mi<strong>to</strong>tic division<br />

continues during sporangial development, which<br />

has not been reported from the other species<br />

studied. The cleavage of the sporangial cy<strong>to</strong>plasm<br />

<strong>to</strong> form spores is accomplished by the<br />

fusion of membranous cleavage vesicles lined by<br />

electron-opaque granules. The vesicles are at first<br />

globose but coalesce and become flattened <strong>to</strong><br />

form cleavage furrows. A three-dimensional network<br />

of cleavage furrows envelops the individual<br />

spore pro<strong>to</strong>plasts, radiating outwards until they<br />

fuse with the sporangial plasma membrane.<br />

After cleavage, the flattened membrane which<br />

bounded the cleavage vesicle persists as the<br />

plasma membrane of the sporangiospore,<br />

whereas the electron-opaque granules make up<br />

part of the spore wall (Fig. 7.7). The columella<br />

is delimited from the rest of the sporangium by<br />

a process similar <strong>to</strong> that which cuts out the<br />

spores. Edelman and Klomparens (1994) noted<br />

that in Z. heterogamus the wall of the sporangium<br />

contains chitin, but the walls of the sporangiospores<br />

and columella do not. They have suggested<br />

that the columella may be a source of<br />

chitinase which causes enzymatic degradation of<br />

the mature sporangial wall whilst not affecting<br />

the walls of the spores or the columella itself.<br />

Fig 7.7 Zygorhynchus heterogamus. Diagrammatic interpretation of ultrastructural transition from cleavage furrows isolating spore<br />

pro<strong>to</strong>plasts (a) <strong>to</strong> post-cleavage spores (b). (a) shows the cleavage furrow membrane (1), an electron-translucent layer (2), a layer of<br />

fusing electron-opaque granules (3), an electron-transparent zone (5) and the electron-translucent matrix of the cleavage furrow<br />

(6). (b) shows the spore plasma membrane which was initially the cleavage furrow membrane (1), an electron-translucent layer (2),<br />

the electron-opaque layer which originated from the fusing granules of the cleavage furrows (3), an additional electron-translucent<br />

layer which had no corresponding layer within the cleavage furrows (4), a uniform zone of electron transparency, similarly placed in<br />

the cleavage furrows (5), and a non-uniform granular matrix separating the spores (6).Redrawn with permission from Edelmann and<br />

Klomparens (1994), Mycologia. ßThe Mycological Society of America.


ZYGOMYCETES: MUCORALES<br />

173<br />

The sporangial wall is sometimes colourless<br />

or yellow, but it often darkens and develops a<br />

spiny surface due <strong>to</strong> the formation of crystals of<br />

calcium oxalate dihydrate (weddellite) beneath<br />

the surface layer (Fig. 7.14c; Jones et al., 1976;<br />

Urbanus et al., 1978; Whitney & Arnott, 1986). In<br />

some species similar crystals may develop on the<br />

sporangiophore. A possible function ascribed <strong>to</strong><br />

these structures is that they form a barrier<br />

against grazing arthropods.<br />

Despite the apparent similarity in sporangial<br />

structure across members of the Mucorales,<br />

spore liberation may be brought about by two<br />

different mechanisms (Ingold & Zoberi, 1963;<br />

Zoberi, 1985). In many of the commonest species<br />

of Mucor (e.g. M. hiemalis), the sporangium wall<br />

dissolves and the sporangium becomes converted<br />

at maturity in<strong>to</strong> a ‘sporangial drop’ adhering <strong>to</strong><br />

the columella. Sporangial walls which dissolve in<br />

this way are said <strong>to</strong> be diffluent. In large<br />

sporangia, for example of M. plasmaticus, M.<br />

mucedo and Phycomyces, the spores are embedded<br />

in mucilage. The sporangial wall does not break<br />

open spontaneously, but the slimy contents<br />

exude when the wall is <strong>to</strong>uched. Such sticky<br />

spore masses are distributed by insects or rain<br />

splash, or by wind after drying. In the second<br />

spore liberation mechanism, the sporangial wall<br />

breaks in<strong>to</strong> pieces, and here air currents or<br />

mechanical agitation readily liberate spores. An<br />

example of this is Mucor plumbeus (Figs. 7.14c,d).<br />

In M. plumbeus the columella terminates in<br />

one or more finger-like or spiny projections<br />

(Fig. 7.14d), and in some Absidia spp. the<br />

columella may also bear a single nipple-like<br />

projection (Fig. 7.17b). In Rhizopus s<strong>to</strong>lonifer the<br />

columella is large, and as the sporangium dries<br />

the columella collapses so that it appears like a<br />

basin balanced at the end of the sporangiophore<br />

(Fig. 7.16d). Associated with these changes in<br />

columella shape, the sporangium wall breaks up<br />

in<strong>to</strong> many fragments and the dry spores can<br />

escape in air currents.<br />

7.2.5 Sexual reproduction<br />

The Mucorales reproduce sexually by a process of<br />

gametangial conjugation resulting in the formation<br />

of zygospores. By strict definition, what we<br />

describe as a zygospore is actually a zygosporangium,<br />

the dark warty ornamentation representing<br />

its outer wall (Benny et al., 2001).<br />

According <strong>to</strong> this definition the zygosporangium<br />

contains a single globose zygospore, sometimes<br />

referred <strong>to</strong> as the zygospore proper. For convenience,<br />

we continue <strong>to</strong> use the term ‘zygospore’<br />

in a wide sense.<br />

Some species are homothallic, zygospores<br />

being formed in cultures derived from a<br />

single sporangiospore (e.g. Rhizopus sexualis,<br />

Syzygites megalocarpus, Zygorhynchus moelleri and<br />

Absidia spinosa). However, the majority of species<br />

are heterothallic and only form zygospores<br />

when compatible strains are mated <strong>to</strong>gether.<br />

It is believed that homothallic species are derived<br />

from heterothallic ances<strong>to</strong>rs (O’Donnell et al.,<br />

2001). There is, in reality, no absolute distinction<br />

between the homothallic and heterothallic<br />

conditions because some species normally<br />

homothallic or heterothallic are ambivalent, i.e.<br />

they can change their mating behaviour under<br />

certain conditions (Schipper & Stalpers, 1980).<br />

Zygospore formation is affected by environmental<br />

conditions, being generally favoured by<br />

darkness (Hesseltine & Rogers, 1987; Schipper,<br />

1987). The effects of temperature are variable. In<br />

Mucor piriformis lower temperatures (0 15°C,<br />

optimum 10°C) favour zygospore formation,<br />

whilst for Choanephora cucurbitarum the optimum<br />

is 20°C (Michailides et al., 1997).<br />

In heterothallic species, if the appropriate<br />

strains are inoculated at opposite sides of a Petri<br />

dish, the mycelia grow out and a line of<br />

zygospores develops where they meet (Fig. 7.8).<br />

The two compatible strains rarely differ from<br />

each other in any obvious morphological or<br />

physiological features, although there may be<br />

slight differences in growth rate and carotenoid<br />

content. Because it was not possible <strong>to</strong> designate<br />

one strain as male and the other as female,<br />

Blakeslee (1906) labelled them (þ) and ( ).<br />

The two compatible strains are said <strong>to</strong> differ in<br />

mating type. The morphological events preceding<br />

zygospore formation are sufficiently similar<br />

<strong>to</strong> allow a general description of the process.<br />

When two compatible strains approach each<br />

other, three reactions can be distinguished.


174 ZYGOMYCOTA<br />

Fig 7.8 Phycomyces blakesleeanus.Maltagarplate5days<br />

after inoculation with a (þ)anda( )strain.Alineof<br />

black zygospores has been produced where the two<br />

mycelia have met.<br />

(1) A ‘telemorphotic reaction’ which involves<br />

the formation of aerial (or occasionally<br />

submerged) swollen hyphal tips. These are<br />

called zygophores or, when they have made<br />

contact with each other, progametangia. They<br />

are often coloured yellow due <strong>to</strong> a high b-<br />

carotene content. (2) A ‘zygotropic reaction’, in<br />

which directed growth of zygophores of (þ) and<br />

( ) mating partners <strong>to</strong>wards each other is<br />

observed. (3) A ‘thigmotropic reaction’, i.e.<br />

a <strong>to</strong>uch response involving the events which<br />

occur after contact of the respective zygophores,<br />

such as gametangial fusion and septation of<br />

the progametangia <strong>to</strong> form gametangia and<br />

suspensors.<br />

Hormonal control of sexual reproduction<br />

The mating process is under the control of<br />

mating hormones (sex hormones, gamones,<br />

pheromones), and the hormones involved are<br />

effective in all members of the Mucorales studied<br />

(see Gooday, 1994; Gooday & Carlile, 1997). Early<br />

evidence of the involvement of pheromones<br />

was the demonstration that in Mucor mucedo the<br />

mating process can be initiated between mycelia<br />

of different mating types separated by a collodion<br />

membrane. The effect of the mating<br />

hormone is <strong>to</strong> switch the vegetative mycelium<br />

from asexual <strong>to</strong> sexual development. Other<br />

effects are the accumulation of carotenoids in<br />

cultures containing both mating types and, in<br />

Phycomyces, of a marked reduction in the growth<br />

rate of the vegetative mycelia as they approach<br />

each other (Drinkard et al., 1982). The mating<br />

hormones have been identified as trisporic acid,<br />

actually a family of structurally related molecules,<br />

and its precursors. Trisporic acid was so<br />

named after Blakeslea trispora (see Fig. 7.27) from<br />

which this substance was first isolated (Austin<br />

et al., 1969; Sutter, 1987). Liquid media inoculated<br />

with a mixture of (þ) and ( ) spores of<br />

B. trispora developed more intense yellow pigmentation<br />

than unmated cultures due <strong>to</strong> a<br />

massive stimulation of b-carotene synthesis.<br />

Trisporic acid itself is derived from b-carotene<br />

(see Fig. 7.9) and is synthesized by collaborative<br />

metabolism of the two different mating type<br />

strains. Each strain has an incomplete enzyme<br />

pathway for the synthesis of trisporic acid so that<br />

intermediates accumulate which can only be<br />

metabolized further by mycelium of the opposite<br />

mating type. As shown in Fig. 7.9, the enzymatic<br />

steps in the conversion of b-carotene (I) <strong>to</strong><br />

4-dihydrotrisporol (III) via retinal (II) are<br />

common <strong>to</strong> both mating types. The (þ) strain<br />

can convert 4-dihydrotrisporol <strong>to</strong> methyl-4-dihydrosporate<br />

(IV), whereas the ( ) strain converts<br />

4-dihydrotrisporol <strong>to</strong> trisporol (V). Thus IV and V<br />

function as two complementary prohormones,<br />

each of which is inactive in its own mycelium


ZYGOMYCETES: MUCORALES<br />

175<br />

but is converted <strong>to</strong> the active hormone trisporic<br />

acid after diffusion in<strong>to</strong> the mycelium of the<br />

complementary strain (Gooday, 1994). Trisporic<br />

acid stimulates further synthesis of b-carotene<br />

and of the two prohormones, leading <strong>to</strong> amplification<br />

of its own synthesis by a ‘cascade’<br />

mechanism. The 15 20-fold enhanced synthesis<br />

of b-carotene upon mating of two compatible<br />

strains of B. trispora holds potential for commercial<br />

production of this substance (Lampila et al.,<br />

1985; Sandmann & Misawa, 2002).<br />

The role of the trisporic acid in inducing<br />

b-carotene synthesis and zygophore formation is<br />

widespread in the Mucorales, having been<br />

characterized in Blakeslea, Phycomyces, Mucor and<br />

even Mortierella (Schimek et al., 2003). It is also<br />

known that trisporic acid is involved in the<br />

sexual response of some homothallic Mucorales<br />

such as Zygorhynchus moelleri, Mucor genevensis and<br />

Syzygites megalocarpus (Lampila et al., 1985). The<br />

common nature of the hormones of homothallic<br />

and heterothallic species could also be inferred<br />

from earlier observations of attempted matings<br />

between such forms, either at the interspecific or<br />

intergeneric level.<br />

Zygophores show directional growth <strong>to</strong>wards<br />

each other in response <strong>to</strong> volatile hormones.<br />

Gooday (1994) has suggested that these may be<br />

the mating type-specific prohormones, methyl-<br />

4-dihydrosporate of the (þ) strain and trisporol of<br />

the ( ) strain.<br />

Thigmotropic reactions<br />

When compatible zygophores make contact, they<br />

become firmly attached <strong>to</strong> each other and<br />

develop in<strong>to</strong> progametangia. In Mucor mucedo<br />

there is evidence that the cell wall chemistry of<br />

the zygophores is distinct from that of the<br />

vegetative mycelium and that the (þ) and ( )<br />

zygophores are bound <strong>to</strong>gether by lectins, i.e.<br />

glycoproteins exhibiting specific binding for<br />

polysaccharides (Jones & Gooday, 1977). In<br />

Fig 7.9 Collaborative biosynthesis of trisporic acid<br />

by cross-feeding of intermediates between (þ)and<br />

( ) mating types of Blakeslea trispora. b-Carotene<br />

(I) is metabolizedby both (þ)and( ) mating-types<br />

via retinal (II) <strong>to</strong> 4-dihydrotrisporol (III).This is<br />

metabolized by (þ)strains<strong>to</strong><br />

4-dehydrosporic acid and its methyl ester (IV)and<br />

by ( ) strains <strong>to</strong> trisporol (V).These are converted<br />

<strong>to</strong> trisporic acid (VI) only after diffusing <strong>to</strong> the ( )<br />

and (þ) strains, respectively. Redrawn from<br />

Gooday (1994), with kind permission of Springer<br />

Science and Business Media.


176 ZYGOMYCOTA<br />

Phycomyces, after arrest of the growth of vegetative<br />

hyphae, certain submerged hyphal tips<br />

develop short branches called ‘knobbly knots’<br />

(Fig. 7.18) which break through the surface of the<br />

agar and become progametangia (O’Donnell<br />

et al., 1976; Sutter, 1987). The progametangia<br />

become tightly appressed and their close contact<br />

is enhanced by the formation of extracellular<br />

fimbriae whose presence appears <strong>to</strong> be essential<br />

for further development in Phycomyces (Yamakazi<br />

& Ootaki, 1996) as well as in other groups of<br />

fungi (see p. 652). Fimbriae may be the lectinbearing<br />

structures. In other Mucorales the<br />

zygophores are aerial and club-shaped. The tip<br />

of each progametangium becomes cut off by a<br />

septum <strong>to</strong> separate a distal multinucleate gametangium<br />

from the subterminal suspensor (see<br />

Fig. 7.10).<br />

The walls separating the two gametangia<br />

break down so that the numerous nuclei from<br />

each cell become surrounded by a common<br />

cy<strong>to</strong>plasm. The fusion cell, or zygote, swells and<br />

develops a dark warty outer layer <strong>to</strong> become the<br />

zygospore.<br />

Cy<strong>to</strong>logy of zygospore formation<br />

There have been numerous accounts of the<br />

cy<strong>to</strong>logy of zygospore formation. Meiosis usually<br />

occurs before zygospore germination, so that the<br />

zygospore can be regarded as a meiosporangium.<br />

Four main types of nuclear behaviour can be<br />

distinguished.<br />

(1) In Mucor hiemalis, Absidia spinosa and some<br />

other species, all the nuclei fuse in pairs within a<br />

few days, then quickly undergo meiosis so that<br />

the mature zygospore contains only haploid<br />

nuclei.<br />

(2) In Rhizopus s<strong>to</strong>lonifer and Absidia glauca,<br />

some of the nuclei entering the zygospore do not<br />

pair, but degenerate. The remainder fuse in<br />

pairs, but meiosis is delayed until germination<br />

of the zygospore.<br />

(3) In Phycomyces blakesleeanus, the haploid<br />

nuclei continue <strong>to</strong> divide mi<strong>to</strong>tically in the<br />

young zygospore and then become associated in<br />

groups, with occasional single nuclei also<br />

present. Before germination some of the nuclei<br />

pair up, and in the germ sporangium diploid<br />

nuclei and also haploid nuclei are found; some of<br />

these may be products of meiosis, others may<br />

represent the scattered solitary nuclei which<br />

failed <strong>to</strong> pair up.<br />

(4) In Syzygites megalocarpus mi<strong>to</strong>tic nuclear<br />

divisions continue in the young zygospore, but<br />

nuclear fusion and meiosis apparently do not<br />

occur. This fungus can therefore be described as<br />

amictic (Burnett, 1965).<br />

The fine structure of zygospore development<br />

has been studied in the homothallic Rhizopus<br />

sexualis (Hawker & Beckett, 1971; Ho & Chen,<br />

1998). Following contact of the tips of the two<br />

zygophores, their walls adhere <strong>to</strong> each other and<br />

become flattened (Figs. 7.10a c) <strong>to</strong> form the<br />

fusion septum. On either side of the fusion<br />

septum, each cell becomes distended <strong>to</strong> form<br />

a progametangium. In each progametangium,<br />

an oblique septum, concave <strong>to</strong> the developing<br />

zygospore, develops by gradual inward extension<br />

mediated by the coalescence of vesicles. When<br />

the septum is complete, it separates the terminal<br />

gametangium from the progametangial base<br />

now called the suspensor. However, cy<strong>to</strong>plasmic<br />

continuity between the suspensor and the<br />

gametangium persists through a series of pores<br />

which probably enable nutrients <strong>to</strong> flow in<strong>to</strong> the<br />

developing zygospore from the surrounding<br />

mycelium. Numerous nuclei congregate on<br />

either side of the fusion septum. It has been<br />

estimated that there may be over 150 nuclei in<br />

a pair of progametangia, but the number<br />

may rise <strong>to</strong> over 300 in a pair of completely<br />

delimited gametangia, reflecting further nuclear<br />

divisions.<br />

The breakdown of the fusion septum is<br />

associated with an accumulation of vesicles in<br />

the vicinity of the dissolving wall. These may<br />

contain wall-degrading enzymes. The fusion<br />

septum is completely dissolved, and once the<br />

cy<strong>to</strong>plasmic contents of the two gametangia are<br />

continuous, the nuclei become arranged in the<br />

periphery of the cy<strong>to</strong>plasm. In R. sexualis, itis<br />

probable that most of the gametangial nuclei<br />

fuse in pairs immediately, and that the fusion<br />

nuclei then quickly divide.<br />

Even before dissolution of the fusion wall is<br />

complete, the primary outer wall of the zygote<br />

thickens, and beneath this original wall, the<br />

warts (which will eventually ornament the wall


ZYGOMYCETES: MUCORALES<br />

177<br />

Fig 7.10 Rhizopus sexualis.(a g) Successive stages in the<br />

formation of zygospores.The fungus is homothallic.<br />

of the mature zygospore) are initiated as widely<br />

separated patches shaped like inverted saucers<br />

(Fig. 7.11c). The cy<strong>to</strong>plasm fills the domes of the<br />

‘saucers’ and also balloons out between them,<br />

enveloped by the plasmalemma. As the zygospore<br />

continues <strong>to</strong> enlarge, the saucers change in<br />

shape and size <strong>to</strong> resemble inverted flower pots<br />

which increase in size by the addition of new<br />

material at their rims until they are contiguous.<br />

From this moment onwards, electron microscopy<br />

fixatives can no longer penetrate, explaining<br />

why cy<strong>to</strong>logical studies of later stages of<br />

zygospore development have proven difficult.<br />

Eventually, the tips of the warts become pushed<br />

through the original primary wall. At least three<br />

wall layers are deposited beneath the original<br />

primary wall (Fig. 7.11g). The darkening of the<br />

wall is probably due <strong>to</strong> the deposition of<br />

melanin. The sculpturing of the zygospore wall<br />

of other members of the Mucorales, as seen by<br />

scanning electron microscopy, shows different<br />

patterns, ranging from circular or conical warts<br />

<strong>to</strong> branched stellate warts. Essentially similar<br />

zygospore development has been reported in<br />

the heterothallic Gilbertella persicaria (O’Donnell<br />

et al., 1977a).<br />

In M. mucedo and P. blakesleeanus, the wall of<br />

the zygospore is rich in sporopollenin (Gooday<br />

et al., 1973; Furch & Gooday, 1978). This substance,<br />

which is also present in the walls of<br />

pollen grains, is extremely resistant <strong>to</strong> degradation<br />

and enables zygospores <strong>to</strong> remain dormant<br />

but undamaged in the soil for long periods.<br />

Sporopollenin is formed by oxidative polymerization<br />

of b-carotene, and this may explain the<br />

high content of this pigment in developing<br />

zygophores. However, b-carotene and sporopollenin<br />

appear <strong>to</strong> be absent from the zygospore<br />

walls of R. sexualis (Hocking, 1963).<br />

Zygospore investment<br />

In Phycomyces and Absidia, the suspensors may<br />

bear appendages which arch over the zygospore.<br />

In Phycomyces the suspensor appendages are black


178 ZYGOMYCOTA<br />

Fig 7.11 Development of the zygospore wall in Rhizopus sexualis (diagrammatic, after Hawker & Beckett,1971). (a) Primary wall<br />

before inflation of zygospore, showing thin electron-dense outer layer, thicker less electron-dense inner one, and scattered<br />

lomasome-like bodies. (b) Blocks of secondary material (wart initials) developing locally on inner surface of primary wall. (c) Wart<br />

initials growing by deposition of secondary material at the rims <strong>to</strong> give saucer-shapedpigmentedmasses. (d) Warts becoming flower<br />

pot shaped by further growth at rims, inner layer of primary wall becoming gelatinous and swollen. Note pockets of cy<strong>to</strong>plasm<br />

between warts. (e) Rims of warts nearly <strong>to</strong>uching, inner layer of primary wall showing stress lines, pockets of cy<strong>to</strong>plasm between<br />

warts much reduced. (f) Edges of warts <strong>to</strong>uching, warts lined with tertiary smoothing layer, outer layer of primary wall <strong>to</strong>rn.<br />

(g) Thick stratified impermeable layer of quaternary material laid down inside smoothing layer, inner gelatinous layer of primary<br />

wall has collapsed as a horny skin enveloping the warts.<br />

and forked, whilst in Absidia they are hyaline<br />

and coiled or curved inwards (see Figs. 7.17, 7.18).<br />

The function of such appendages is unknown;<br />

possibly they assist in attaching zygospores <strong>to</strong><br />

passing animals. The forked appendage tips<br />

of Phycomyces bear a drop of liquid, and they<br />

have been interpreted as hydathodes (i.e. watersecreting<br />

structures). In the homothallic species<br />

A. spinosa the appendages arise on only one<br />

suspensor.<br />

Mating behaviour<br />

Analysis of the results of crosses involving several<br />

genes suggest that there is a single mating type<br />

locus with two alternative alleles, (þ) and ( ),<br />

which segregate at meiosis. However, no DNA<br />

sequences of this locus have as yet been<br />

published, and there are also a number of<br />

anomalous results for which a full cy<strong>to</strong>logical<br />

explanation is still awaited.<br />

Hybridization experiments have been<br />

conducted between different species and genera<br />

of Mucorales, and in some cases imperfect<br />

zygospores are formed. Attempted copulation<br />

has also been observed between homothallic and<br />

heterothallic strains. An unusual type of mating<br />

behaviour has been discovered in Mucor pusillus<br />

which is predominantly heterothallic but in<br />

which homothallic strains are known. It has<br />

been possible <strong>to</strong> induce a (þ) strain <strong>to</strong> mutate<br />

<strong>to</strong> a ( ) strain, and also <strong>to</strong> a homothallic strain<br />

by g-irradiation (Nielsen, 1978).<br />

7.2.6 Zygospore germination<br />

After a resting period the zygospore may germinate<br />

by developing a germ sporangium which<br />

resembles an ordinary sporangium and contains<br />

sporangiospores of the normal type. In some<br />

cases vegetative mycelium develops from the<br />

germinating zygospore. The conditions for zygospore<br />

germination are, in many cases, imperfectly<br />

known, but a pro<strong>to</strong>col for germination<br />

has been established for Phycomyces blakesleeanus<br />

(Eslava & Alvarez, 1987). Mature zygospores


ZYGOMYCETES: MUCORALES<br />

179<br />

collected from 6-week-old cultures and placed on<br />

moist filter paper at 22°C under alternating<br />

light/dark illumination will germinate after<br />

about 8 days, reaching maximum germination<br />

after a further 8 10 days. In Mucor piriformis, the<br />

germination rate is highest in fresh zygospores<br />

(Guo & Michailides, 1998), a vigorous germ tube<br />

emerging through one of the suspensors or<br />

through a crack in the zygospore wall. The<br />

germ tube may continue development as mycelium<br />

or grow in<strong>to</strong> the air and form a germ<br />

sporangium at the tips of single or branched<br />

sporangiophores.<br />

Mating-types represented in germ sporangia<br />

The distribution of mating types amongst the<br />

germ spores which are present in germ sporangia<br />

falls in<strong>to</strong> three categories.<br />

1. Pure germinations in which all the spores<br />

are homothallic, e.g. in Mucor genevensis,<br />

Zygorhynchus dangeardi and Syzygites megalocarpus.<br />

2. Pure germinations in which all sporangiospores<br />

are of one mating type, i.e. all (þ)<br />

or all ( ). Mucor mucedo, M. hiemalis and<br />

P. blakesleeanus generally behave in this way. In<br />

P. blakesleeanus, the analysis of progeny from<br />

crosses involving up <strong>to</strong> four unlinked fac<strong>to</strong>rs<br />

which included mating type were best explained<br />

on the basis of the survival of a single diploid<br />

nucleus from the thousands which are present<br />

in the young zygospore (Cerdá-Olmedo, 1975;<br />

Eslava et al., 1975a,b). This single diploid nucleus<br />

undergoes meiosis and one or more of the<br />

resultant nuclei divide mi<strong>to</strong>tically <strong>to</strong> provide<br />

nuclei for the germ sporangium (Fig. 7.12).<br />

Occasionally two or three diploid nuclei may<br />

survive and undergo meiosis. In some germ<br />

sporangia heterokaryotic spores are present. If<br />

these are heterokaryotic for mating type, the<br />

mycelium which develops from them may be<br />

abnormal and ‘neuter’, i.e. it is unable <strong>to</strong> mate<br />

with (þ) as well as ( ) strains.<br />

3. Mixed germinations. In Phycomyces nitens,<br />

the same germ sporangium sometimes contains<br />

(þ), ( ) and homothallic (i.e. self-fertile) spores.<br />

The finding that diploid nuclei enter the<br />

germ sporangia may be the explanation for the<br />

presence of homothallic spores which should<br />

properly be described as secondarily<br />

homothallic. Mixed germinations have also<br />

been reported by Gauger (1961) for Rhizopus<br />

s<strong>to</strong>lonifer in which both (þ) and ( ) spores were<br />

present in some germ sporangia, whereas others<br />

contained spores of either mating type. For this<br />

type of mixed germination <strong>to</strong> occur, it would<br />

be necessary only <strong>to</strong> postulate the survival of<br />

more than one meiotic product so that both<br />

mating types are represented in the sporangium.<br />

‘Neuter’ spores were found in some germ<br />

sporangia. Thus, in Choanephora cucurbitarum<br />

mixed germinations have been reported in<br />

which the majority (usually all) of the germ<br />

sporangia contained only either (þ) or( ) spores,<br />

but a low proportion gave heterokaryotic spores<br />

of mating-type (þ/ ). A characteristic feature of<br />

C. cucurbitarum cultures derived from heterokaryotic<br />

(þ/ ) germ spores or fusion of (þ) with ( )<br />

pro<strong>to</strong>plasts is that they produce azygospores<br />

(Yu & Ko, 1996, 1999; see below).<br />

7.2.7 Azygospores<br />

In some Mucorales, if gametangial copulation<br />

fails <strong>to</strong> take place normally, one or both<br />

gametangia may give rise parthenogenetically<br />

<strong>to</strong> a structure morphologically similar <strong>to</strong> the<br />

zygospore, termed an azygospore (azygosporangium).<br />

Azygospores therefore usually appear as<br />

warty spherical structures borne on a single<br />

suspensor-like cell, or occasionally on a sporangiophore.<br />

They are formed regularly in cultures<br />

of Mucor bainieri and M. azygospora (Fig. 7.13), both<br />

of which are obligately azygosporic and do not<br />

form true zygospores (Benjamin & Mehrotra,<br />

1963), and they have also been reported in<br />

Rhizopus azygosporus (Yuan & Yong, 1984). The<br />

development of azygospores of M. azygospora<br />

resembles that of normal zygospores in other<br />

Mucorales (O’Donnell et al., 1977b; Ginman &<br />

Young, 1989). Azygospore formation may occur<br />

in intergeneric and interspecific crosses, for<br />

example in crosses between a (þ) strain of<br />

Gilbertella persicaria and a ( ) strain of Rhizopus<br />

s<strong>to</strong>lonifer (O’Donnell et al., 1977c) and between<br />

different species of Rhizopus (Schipper, 1987).<br />

Azygospore development has also been seen<br />

in intraspecific crosses, e.g. in certain isolates<br />

of M. hiemalis (Gauger, 1966, 1975). These azygosporic<br />

isolates of M. hiemalis were derived from


180 ZYGOMYCOTA<br />

Fig 7.12 Life cycle of Phycomyces blakesleeanus (diagrammatic and not <strong>to</strong> scale). From a coenocytic haploid mycelium of either<br />

mating type (þ)or( ), sporangiophores develop. Sporangia are columellate and contain numerous sporangiospores which, in<br />

P. blakesleeanus, are multinucleate.When hyphae of both mating types meet, sexual reproduction is initiated by the formation of<br />

knobbly zygophores which develop in<strong>to</strong> progametangia. Each progametangium divides in<strong>to</strong> a gametangium and a suspensor, the<br />

latter ornamented by black forked appendages. Plasmogamy (P) occurs by lysis of the wall separating the two multinucleate<br />

gametangia.This is followed by mass karyogamy (K), but only one of the numerous diploid fusion nuclei seems <strong>to</strong> undergo<br />

meiosis (M), and only one of the resulting tetrad nuclei survives in the zygospore during dormancy, so that the sporangiospores<br />

in the germ sporangium are usually of either one or the other mating type.Open and closed circles represent haploid nuclei of<br />

opposite mating type; diploid nuclei are larger and half-filled.<br />

spores of germ sporangia developed from normal<br />

zygospores. If the azygosporic strains are subcultured,<br />

either from single sporangiospores or by<br />

mass transfer, they show a tendency <strong>to</strong> ‘break<br />

down’ <strong>to</strong> strains of (þ) or ( ) mating type of<br />

normal appearance. It seems that azygosporic<br />

strains of M. hiemalis are typically diploid and<br />

heterozygous for mating type, i.e. the diploid<br />

nucleus carries both (þ) and ( ) mating type<br />

alleles. The breakdown <strong>to</strong> the normal (þ) or( )<br />

mating type condition may be brought about<br />

by somatic (i.e. non-meiotic) reduction leading <strong>to</strong><br />

aneuploid intermediates, and finally <strong>to</strong> haploids.<br />

The germination of azygospores is unknown.<br />

7.3 Examples of Mucorales<br />

As mentioned before, the traditional family<br />

classification within the Mucorales is artificial<br />

(see Benny et al., 2001; Tanabe et al., 2004),<br />

and we use it here solely for convenience of<br />

presentation.<br />

7.3.1 Mucoraceae<br />

Mucor<br />

About 50 species of Mucor are currently known<br />

(Kirk et al., 2001). The genus is cosmopolitan, with


EXAMPLES OF MUCORALES<br />

181<br />

Fig 7.13 Azygospore of Mucor azygospora.Originalimage<br />

kindly provided byT.W.K.Young.<br />

many species being widespread in soil or<br />

on substrates in contact with soil. Most species<br />

are mesophilic (growing at 10 40°C with an<br />

optimum 20 35°C), but some, e.g. M. miehei or<br />

M. pusillus (sometimes classified as species of<br />

Rhizomucor; see Mouchacca, 1997, 2000) are<br />

thermophilic, with a minimum growth temperature<br />

of about 20°C and a maximum extending up<br />

<strong>to</strong> 60°C (Cooney & Emerson, 1964; Maheshwari<br />

et al., 2000). Mucor indicus and M. circinelloides are<br />

used as starters in food processing <strong>to</strong> break down<br />

starchy polysaccharides in rice, cassava and<br />

sorghum, releasing simple sugars for the<br />

preparation of fermented foods or alcohol<br />

production (Hesseltine, 1991).<br />

Most species of Mucor grow rapidly on agar<br />

at room temperature, filling a Petri dish in<br />

2 3 days with their coarse aerial mycelium.<br />

When incubated in liquid culture under semianaerobic<br />

conditions, several species grow in<br />

a yeast-like state. The ability <strong>to</strong> switch between<br />

the yeast-like and filamen<strong>to</strong>us state is termed<br />

dimorphism, a phenomenon which has been<br />

studied in greatest detail in M. rouxii (see<br />

Fig. 7.14), but also occurs in M. circinelloides,<br />

M. fragilis, M. hiemalis, M. lusitanicus and in other<br />

Mucorales (Orlowski, 1991, 1995). Sporangia are<br />

globose and borne on branched and unbranched<br />

sporangiophores growing in<strong>to</strong> the air. The<br />

columella is large and typically elongated<br />

(Figs. 7.2 and 7.3). Zygospores are rarely formed<br />

in agar culture because most species are heterothallic.<br />

Amongst the most common species from<br />

soil are M. hiemalis, M. racemosus and M. spinosus<br />

(Domsch et al., 1980). Several species of Mucor,<br />

e.g. M. mucedo, fruit on dung (Ellis & Ellis, 1998;<br />

Richardson & Watling, 1997), and they are the<br />

earliest fungi <strong>to</strong> appear in the succession of<br />

fungal fruit bodies on this substrate (Dix &<br />

Webster, 1995). The sporangiospores of coprophilous<br />

Mucor spp. survive digestion by herbivorous<br />

mammals.<br />

A few species of Mucor are human pathogens.<br />

The term mucormycosis, however, usually refers<br />

<strong>to</strong> conditions caused by Mucorales generally<br />

rather than the genus Mucor (Rinaldi, 1989;<br />

Eucker et al., 2001) because it is not possible <strong>to</strong><br />

identify species by the microscopic appearance<br />

of their coenocytic mycelium within diseased<br />

tissue. Diagnosis is dependent on the isolation<br />

and identification of the suspected pathogen in<br />

culture, sometimes post mortem. By these means,<br />

several ubiqui<strong>to</strong>us species of Mucor have been<br />

associated with disease symp<strong>to</strong>ms, including<br />

M. circinelloides, M. hiemalis and M. racemosus.<br />

Infections are opportunistic, derived from sporangiospores<br />

present in the soil or air, and are<br />

usually associated with patients suffering from<br />

other diseases such as diabetes, leukaemia,<br />

AIDS and post-operative conditions. There are<br />

no records of person-<strong>to</strong>-person transmission.<br />

Mucormycoses are serious, even fatal in immunocompromised<br />

patients, although some can be<br />

successfully treated by surgery and antibiotics<br />

such as amphotericin B (Kwon-Chung & Bennett,<br />

1992).<br />

Schipper (1978) has given a key <strong>to</strong> 49 species<br />

of Mucor, and Watanabe (1994) has described the<br />

six homothallic and two azygosporic species.


182 ZYGOMYCOTA<br />

Fig 7.14 Mucor racemosus (a,b) and M. plumbeus (c,d). (a) Tip of a sporangiophore which has formed a sporangium.The columella<br />

(arrow) is visible. (b) Lower region of sporangiophores showing intercalary thick-walled chlamydospores which are typical of the<br />

species. (c) Sporangium with a spiny surface of calcium oxalate crystals. (d) Exposed columellae with finger-like projections. (a,b) <strong>to</strong><br />

same scale; (c,d) <strong>to</strong> same scale.<br />

Zygorhynchus<br />

There are about six species, mostly reported from<br />

soil, often from considerable depth (Hesseltine<br />

et al., 1959). All species are homothallic and,<br />

unusually, form heterogametangic zygospores<br />

(Fig. 7.15). The sporangiophores are commonly<br />

branched and the columella is often broader than<br />

high. The most frequently encountered species is<br />

Z. moelleri, which has been isolated worldwide<br />

from a range of soils and from the rhizosphere of<br />

numerous plants (Domsch et al., 1980). Most<br />

species are mesophilic, but Z. psychrophilus forms<br />

zygospores readily at 5°C. Sporangium development<br />

in Z. heterogamus has been studied by<br />

Edelmann and Klomparens (1994) (see p. 171).<br />

Zygospore development and structure in several<br />

species of Zygorhynchus have been described by<br />

O’Donnell et al. (1978a). The warts on the outside<br />

of the zygosporangia often appear as interlocking,<br />

starfish-like pointed thickenings. The inner<br />

wall of the zygosporangium is ornamented by a<br />

network of ridges and grooves radiating from<br />

centres corresponding <strong>to</strong> the points of the warts.<br />

The outer wall of the zygospore proper, lying<br />

within the zygosporangium, is similarly<br />

ornamented by a pattern of radiating grooves<br />

and ridges which are a template of the lining<br />

of the zygosporangial wall.<br />

Rhizopus<br />

There are about 10 species which grow in soil<br />

(Domsch et al., 1980) and on fruits, other foods<br />

and all kinds of decaying materials. Rhizopus spp.<br />

also occur frequently as labora<strong>to</strong>ry contaminants.<br />

Rhizopus s<strong>to</strong>lonifer (syn. R. nigricans) grows<br />

rapidly. It is often found on ripe fruits, especially<br />

if these are incubated in a moist atmosphere (see<br />

Plate 3d). Characteristic features of Rhizopus are<br />

the presence of rhizoids at the base of the<br />

sporangiophores (which may grow in clusters),<br />

and the s<strong>to</strong>loniferous habit (Fig. 7.16). An aerial<br />

hypha grows out, and where it <strong>to</strong>uches on the<br />

substratum it bears rhizoids and sporangiophores.<br />

Growth in this manner is repeated. The<br />

sporangium wall is brittle and the sporangiospores<br />

are dry and wind-dispersed. Some species<br />

of Rhizopus, e.g. R. oryzae, R. microsporus and its<br />

allies, are used as starters in ragi fermentations<br />

of rice (Hesseltine, 1991). Several species<br />

(R. arrhizus, R. microsporus, R. rhizopodiformis) are


EXAMPLES OF MUCORALES<br />

183<br />

Fig 7.15 Zygorhynchus moelleri.<br />

(a) Zygospore and sporangium.<br />

(b) Young sporangiophores.<br />

(c) Dehisced sporangia. (d g) Stages<br />

in zygospore formation. Note that<br />

the fungus is homothallic and that the<br />

suspensors are unequal.<br />

human pathogens associated with mucormycosis<br />

(Rinaldi, 1989). Most species of Rhizopus are<br />

heterothallic, but R. sexualis is homothallic and<br />

forms zygospores freely within 2 days in the<br />

labora<strong>to</strong>ry (see Fig. 7.10).<br />

Rhizopus microsporus causes rice seedling<br />

blight in which root growth is strongly impaired<br />

by a <strong>to</strong>xin, rhizoxin, excreted by the soil-borne<br />

pathogen. The <strong>to</strong>xin binds <strong>to</strong> b-tubulin, thereby<br />

interfering with mi<strong>to</strong>sis. Intriguingly, it is<br />

synthesized not by R. microsporus but by bacteria<br />

(Burkholderia spp.) living endosymbiotically<br />

within the cy<strong>to</strong>plasm of Rhizopus hyphae<br />

(Partida-Martinez & Hertweck, 2005). Bacterial<br />

endosymbionts have been reported only rarely<br />

from fungi, e.g. in the zygomycete Geosiphon<br />

pyriforme (p. 221), or within hyphae of the<br />

ascomycete Morchella elata (p. 427) and the<br />

basidiomycete Laccaria bicolor (p. 552).<br />

Absidia<br />

There are some 20 species growing in soil<br />

(Domsch et al., 1980). Characteristic features are<br />

pear-shaped sporangia arising in partial whorls<br />

along s<strong>to</strong>lon-like branches which produce<br />

rhizoids at intervals but not opposite the


184 ZYGOMYCOTA<br />

that many workers confused the two and much<br />

of the early literature on P. nitens probably refers<br />

<strong>to</strong> P. blakesleeanus. Neither species is particularly<br />

common, but likely substrata are fatty products<br />

and empty oil casks. Bread, dung and decaying<br />

hops are other recorded substrata. The zygospores<br />

are unusual in that they are overarched<br />

by black, forked suspensor appendages (see<br />

Fig. 7.18; O’Donnell et al., 1976, 1978b). Great<br />

interest has been focused on the development<br />

and sensory perception of the spectacularly<br />

large sporangiophore, especially <strong>to</strong> light (see<br />

pp. 169 171), and on the genetics of Phycomyces<br />

(Eslava & Alvarez, 1987). The genus has been<br />

classified in the family Phycomycetaceae by<br />

Benny et al. (2001).<br />

Fig 7.16 Rhizopus s<strong>to</strong>lonifer. (a) Habit sketch, showing<br />

s<strong>to</strong>lon-like branches which develop rhizoids and tufts of<br />

sporangiophores. (b) Two sporangiophores showing basal<br />

rhizoids. (c) Dehisced sporangium showing the columella<br />

with attached spores. (d) Invaginated columella.<br />

sporangiophores. The zygospores are surrounded<br />

by curved unbranched suspensor appendages<br />

which may arise from either or both suspensors<br />

(Fig. 7.17). Most species are heterothallic but<br />

A. spinosa is homothallic. Absidia glauca and<br />

A. spinosa are amongst the most commonly<br />

isolated species. Absidia corymbifera is a human<br />

pathogen.<br />

Phycomyces<br />

The two best-known species are P. blakesleeanus<br />

and P. nitens. The sporangiospores of P. nitens are<br />

larger than those of P. blakesleeanus, but it is likely<br />

Syzygites<br />

Syzygites megalocarpus (¼ Sporodinia grandis; see<br />

Hesseltine, 1957) is found on the decaying<br />

basidiocarps of various <strong>to</strong>ads<strong>to</strong>ols, especially<br />

Boletus, Lactarius and Russula. It grows readily in<br />

culture and is homothallic. Probably because of<br />

this fact it was the first member of the Mucorales<br />

for which sexual reproduction was described in<br />

detail (Davis, 1967). The sporangiophores are<br />

dicho<strong>to</strong>mous and bear thin-walled sporangia<br />

(Fig. 7.19a). In culture, light is essential for the<br />

development of sporangiophores but has no<br />

effect on zygospore formation. Lowering of the<br />

osmotic potential markedly stimulates zygospore<br />

development and this is possibly significant<br />

ecologically in that the drying out of basidiocarp<br />

tissue on which the fungus grows might induce<br />

the development of the zygosporic (resting) state<br />

(Kaplan & Goos, 1982).<br />

7.3.2 Pilobolaceae<br />

There are two common genera, Pilobolus and<br />

Pilaira, which grow on the dung of herbivores.<br />

Both genera have evolved mechanisms <strong>to</strong> ensure<br />

that their sporangia escape from the vicinity of<br />

the dung patch on which they were produced. In<br />

Pilobolus the sporangiophore is swollen and<br />

the sporangium is shot away violently by a jet<br />

of liquid, whilst in Pilaira the sporangiophore<br />

is elongated and the sporangium becomes<br />

converted in<strong>to</strong> a sporangial drop, breaking off


EXAMPLES OF MUCORALES<br />

185<br />

Fig 7.17 Absidia glauca. (a) Habit showing whorls of<br />

pear-shaped sporangia. (b) Intact and dehisced sporangia.<br />

Note the single pointed projection on certain columellae.<br />

(c) Zygospore showing the arching suspensor appendages.<br />

upon contact with an object. The sporangia are<br />

black and melanized, presumably as a protection<br />

against UV irradiation. Grove (1934) has written<br />

a monograph of the family which has s<strong>to</strong>od<br />

the test of time.<br />

Pilobolus<br />

The generic name means literally the ‘hat<br />

thrower’, referring <strong>to</strong> the sporangial discharge<br />

mechanism. If fresh herbivore (e.g. rabbit,<br />

sheep, deer, horse) dung is incubated in light,<br />

the characteristic bulbous sporangiophores of<br />

Pilobolus appear after a preliminary phase of<br />

fruiting of Mucor which may last for 4 7 days<br />

(Fig. 7.20, Plate 3f). Nine Pilobolus spp. previously<br />

recognized have been reduced <strong>to</strong> five, but<br />

including a number of varieties (Hu et al., 1989).<br />

Common species are P. crystallinus, P. kleinii (¼<br />

P. crystallinus var. kleinii), and P. umbonatus (¼<br />

P. roridus var. umbonatus). Here we adopt the<br />

nomenclature proposed by Grove (1934). As far as<br />

is known all members of the Pilobolaceae are<br />

heterothallic.<br />

A full account of the development and<br />

discharge of the sporangium has been given by<br />

Buller (1934) and Ingold (1971). Discharged<br />

sporangia of Pilobolus become attached <strong>to</strong> vegetation<br />

surrounding the dung on which they were<br />

produced. When the vegetation is eaten by a<br />

herbivore, the spores are released in<strong>to</strong> the gut.<br />

In the voided faeces, the spores germinate <strong>to</strong><br />

form a mycelium. After about 4 days, the<br />

mycelium near the surface of the dung pellet<br />

forms trophocysts, swollen hyphal segments<br />

coloured yellow by carotenoids (Fig. 7.20).<br />

Sporangiophores develop from the trophocysts<br />

in a regular daily sequence, and the stage of<br />

development can be correlated with the time of<br />

day. During the late afternoon the sporangiophore<br />

grows away from the trophocyst <strong>to</strong>wards<br />

the light and during the night its tip enlarges <strong>to</strong><br />

become the sporangium. The swelling of the<br />

subsporangial vesicle takes place mainly between<br />

midnight and the early morning. Young sporangiophores<br />

are highly pho<strong>to</strong>tropic even before<br />

their sporangia are differentiated, and the<br />

clear tip of the developing sporangiophore is


186 ZYGOMYCOTA<br />

Fig 7.18 Phycomyces blakesleeanus.Stagesin<br />

zygospore formation.The fungus is<br />

heterothallic. (a) Zygophores consisting of<br />

knobbly hyphal branch tips which become<br />

closely appressed. (b) Paired club-shaped<br />

progametangia which develop from the<br />

appressed zygophores. (c) Septation of the<br />

progametangia <strong>to</strong> form terminal gametangia<br />

and subterminal suspensors. Appendages<br />

are developing on the suspensor <strong>to</strong> the right.<br />

(d) Young zygospore overarched by<br />

dicho<strong>to</strong>mous suspensor appendages.<br />

the sensitive region. Despite the bright yellow<br />

carotenoid deposits in the trophocysts and young<br />

sporangiophores, studies of the pho<strong>to</strong>tropic<br />

response <strong>to</strong> light of various wavelengths suggest<br />

that the pho<strong>to</strong>recep<strong>to</strong>r in the sporangiophore is<br />

more likely <strong>to</strong> be a flavonoid than a carotenoid<br />

(Page & Curry, 1966). Fully developed sporangiophores<br />

are also highly pho<strong>to</strong>tropic. Light<br />

projected along the axis of the sporangiophore<br />

is brought <strong>to</strong> a focus at a point beneath the<br />

swollen vesicle termed the ocellus. In this region,<br />

there is an accumulation of carotenoid-rich<br />

cy<strong>to</strong>plasm which glows orange when illuminated<br />

(Plate 3f). When light falls asymmetrically on<strong>to</strong><br />

the sporangiophore, it is focused on<strong>to</strong> the back of<br />

the subsporangial vesicle near its base, and some<br />

stimulus is probably transmitted <strong>to</strong> the cylindrical<br />

part of the sporangiophore, resulting in<br />

more rapid growth of the wall facing away from<br />

the light. Curvature of the whole sporangiophore<br />

thus occurs until it is again orientated parallel<br />

<strong>to</strong> the incident light (see Fig. 7.21).<br />

The structure of the sporangium differs in a<br />

number of ways from that of the Mucoraceae.<br />

The sporangium is hemispherical, and its wall is<br />

dark black, shiny, <strong>to</strong>ugh and unwettable. At the<br />

base of the sporangium is a conical columella,<br />

which is separated from the spores by a pad of<br />

mucilage. During late morning the sporangium<br />

cracks open by a suture running around the<br />

base, just above the columella. The spores are<br />

prevented from escaping by the mucilaginous<br />

pad which protrudes through the crack in<br />

the sporangium wall as a ring of mucilage<br />

(Figs. 7.20e,f). The subsporangial vesicle is<br />

turgid, and the osmotic pressure of the liquid<br />

has been estimated <strong>to</strong> be around 5.5 bars (Buller,<br />

1934). Drops of liquid decorate the outside of


EXAMPLES OF MUCORALES<br />

187<br />

Fig 7.19 Syzygites megalocarpus.<br />

(a) Sporangiophore. (b) Germinating<br />

zygospore.The fungus is homothallic.<br />

Fig 7.20 Asexual reproduction in Pilobolus<br />

kleinii. (a) Developing trophocyst which is<br />

becoming distended by carotenoid-rich<br />

cy<strong>to</strong>plasm. (b) Trophocyst with immature<br />

sporangiophore.The clear tip of the<br />

sporangiophore is light-sensitive.<br />

(c) Trophocyst bearing a developing<br />

sporangium.The upper part of the sporangium<br />

is beginning <strong>to</strong> darken.Globules of liquid<br />

accumulate on the sporangiophore surface<br />

(9.00 p.m.). (d) Trophocyst with sporangium<br />

which has not yet dehisced (9.00 a.m.).<br />

The arrow (o) points <strong>to</strong> a carotenoid-rich band<br />

of cy<strong>to</strong>plasm called the ocellus.<br />

(e) Sporangiophore bearing a<br />

sporangium which has dehisced near its base.<br />

Spores have extruded and are held in place by<br />

a ring of mucilage (11.30 a.m.). (f) Sporangium<br />

showing dehiscence line at its base<br />

(d). (g) Discharged sporangium surrounded<br />

by dried-out vesicular sap.The spores are<br />

enclosed in mucilage. (a e) <strong>to</strong> same scale;<br />

(f,g) <strong>to</strong> same scale.


188 ZYGOMYCOTA<br />

the sporangiophore of Pilobolus, as they do in<br />

many zygomycetes. Eventually, usually about<br />

midday, the sporangial vesicle explodes at a line<br />

of weakness just beneath the columella. Due <strong>to</strong><br />

the elasticity of the vesicle wall the liquid<br />

contents are squirted out, projecting the entire<br />

sporangium forward in the direction of the light.<br />

Pho<strong>to</strong>graphs of the jet show that it is at first<br />

cylindrical but eventually breaks up in<strong>to</strong> fine<br />

droplets (Fig. 7.22c; Page, 1964). In P. kleinii, the<br />

velocity of projection varies between wide limits<br />

of 4.7 27.5 m s 1 with a mean of 10.8 m s 1<br />

(Page & Kennedy, 1964). The sporangia can be<br />

projected vertically upwards for as much as<br />

2 m and horizontally for up <strong>to</strong> 2.5 m. On striking<br />

a grass blade or other herbage, the sporangium<br />

becomes attached in such a way that the<br />

mucilaginous ring adheres <strong>to</strong> it, with the black<br />

sporangium wall facing outwards. Buller (1934)<br />

has suggested that the projectile contains a<br />

drop of liquid attached <strong>to</strong> the sporangium<br />

(Fig. 7.22a). When the projectile strikes an<br />

object the liquid flows around the sporangium,<br />

but because the sporangium wall is hydrophobic<br />

and the base of the sporangium is surrounded<br />

by the wettable mucilaginous ring, the sporangium<br />

turns round in the liquid so that its wall<br />

faces outwards (Fig. 7.22b). The non-wettable<br />

nature of the sporangial wall may be related <strong>to</strong><br />

the presence on its surface of hollow, blunttipped<br />

spines and crystals as seen by electron<br />

microscopy (Bland & Charles, 1972). As the<br />

mucilage dries, the sporangium becomes<br />

cemented on<strong>to</strong> the surface which it struck. The<br />

spores of Pilobolus are released only after the<br />

sporangium has been ingested by an animal.<br />

They survive gut passage and are voided with<br />

the faeces. A film featuring the life cycle of<br />

Pilobolus has been made (Webster & Hard, 1999).<br />

An unexpected consequence of the attachment<br />

of Pilobolus sporangia <strong>to</strong> herbage is that the<br />

sporangia may act as vec<strong>to</strong>rs for parasitic<br />

nema<strong>to</strong>des such as Dictyocaulus spp., which multiply<br />

on dung and, when ingested, cause lungworm<br />

disease in sheep, cattle and some wild<br />

mammals.<br />

The physiology of Pilobolus shows a number<br />

of interesting features possibly related <strong>to</strong> its<br />

coprophilous habit. Spores germinate best above<br />

Fig 7.21 Pilobolus kleinii. Diagrammatic L.S. of sporangiophore<br />

showing the path of light rays falling parallel <strong>to</strong> the axis of the<br />

sporangiophore which are brought <strong>to</strong> a focus beneath the<br />

subsporangial vesicle.The sporangiophore illustrated is<br />

orientated symmetrically with respect <strong>to</strong> the incident light.<br />

Note the mucilaginous ring extruded through the sporangial<br />

wall at its base (after Buller,1934).<br />

pH 6.5, and can be induced <strong>to</strong> germinate by<br />

treatment with alkaline pancreatin. Germination<br />

can also be triggered by hexoses such as<br />

glucose and mannose. Mycelial growth occurs<br />

over a wide range of temperatures, with optimum<br />

temperatures at 25 35°C. Growth on<br />

synthetic media with asparagine and acetic acid


EXAMPLES OF MUCORALES<br />

189<br />

Fig 7.2 2 Projectiles of Pilobolus (diagrammatic). (a) Sporangium with adherent drop of sporangiophore sap about <strong>to</strong> strike an<br />

obstacle. (b) Sporangium after striking the obstacle.The sporangiophore sap has flowed round the sporangium which has turned<br />

outwards so that the mucilage ring adheres <strong>to</strong> the surface of the obstacle (after Buller,1934). (c) Sporangiophore releasing a<br />

sporangium. Note the jet of liquid and the bending of the narrow base of the sporangiophore under the recoil of the discharge<br />

(after Page,1964).<br />

as nutrients is stimulated by the addition<br />

of thiamine, haemin and coprogen, an organoiron<br />

compound produced by various fungi<br />

and bacteria (Hesseltine et al., 1953; Page, 1960;<br />

Levetin & Caroselli, 1976). Sporangium formation<br />

is stimulated by ammonia, and in dual<br />

cultures Mucor plumbeus may release sufficient<br />

gaseous ammonia <strong>to</strong> induce asexual reproduction<br />

in Pilobolus spp. (Page, 1959, 1960).<br />

Pilaira<br />

Pilaira (Fig. 7.23) also appears early in the<br />

succession of coprophilous fungi, i.e. the order<br />

in which their fruit bodies appear on herbivore<br />

dung incubated under moist conditions. It has<br />

not been found in the tropics (Kirk, 1993). The<br />

structure of the melanized sporangium closely<br />

resembles that of Pilobolus in that the spores are<br />

separated from the columella by a mucilaginous<br />

ring which extrudes from the base of the<br />

sporangium. There are, however, no trophocysts<br />

or subsporangial vesicles, and sporangial release<br />

is non-violent. The cylindrical sporangiophores<br />

are pho<strong>to</strong>tropic, and when mature, especially<br />

under moist conditions, they elongate rapidly<br />

<strong>to</strong> a length of several centimetres (H. J. Fletcher,<br />

1969, 1973). Their development essentially<br />

resembles that of Phycomyces. In a moist atmosphere,<br />

the mucilaginous ring may absorb water<br />

and swell considerably so that a large sporangial<br />

drop is formed (Ingold & Zoberi, 1963). When the<br />

mucilaginous ring at the base of the sporangium


190 ZYGOMYCOTA<br />

makes contact with adjacent herbage it becomes<br />

firmly attached <strong>to</strong> it. The sporangium slips off its<br />

columella and dries down on<strong>to</strong> the herbage.<br />

Pilaira anomala forms zygospores resembling<br />

those of Pilobolus. On germination, a germ<br />

sporangium is produced (Fig. 7.23g).<br />

Nutritional studies on P. anomala indicate a<br />

preference for NH 4 þ and urea over NH 3 or<br />

asparagine as nitrogen sources, and a biotin<br />

and thiamine requirement, but no requirement<br />

of haem compounds for either growth or fruiting,<br />

in contrast <strong>to</strong> the related genus Pilobolus. In<br />

culture, of the simple carbon sources tested only<br />

glucose and fruc<strong>to</strong>se supported good growth,<br />

and there was no evidence of enzymatic ability <strong>to</strong><br />

degrade starch, cellulose, pectin or proteins<br />

(Wood & Cooke, 1987). These nutritional characteristics<br />

support the idea that P. anomala is a<br />

typical ruderal fungus, its activities constrained<br />

by the availability of simple soluble nutrients<br />

which would be depleted rapidly in decomposing<br />

dung.<br />

7.3.3 Thamnidiaceae<br />

In this family two kinds of asexual reproductive<br />

structure are found, namely columellate sporangia<br />

of the Mucor type and smaller, few-spored,<br />

usually non-columellate sporangia termed sporangiola,<br />

which are often borne in whorls or at<br />

the tips of branches. The branches bearing<br />

the sporangiola may be borne laterally on the<br />

columellate sporangiophores or may arise separately.<br />

In some cases the branch system bearing<br />

the sporangiola is terminated by a spine. Benny<br />

et al. (2001) have recognized 10 genera but we<br />

shall consider only Thamnidium.<br />

Fig 7.23 Pilaira anomala. (a) Sporangiophore<br />

from rabbit dung showing rupture of the<br />

sporangial wall at the base of the sporangium.<br />

(b) Sporangium with extruded mucilage ring<br />

adhering <strong>to</strong> an adjacent hypha. (c) Columella<br />

after sporangium has been detached.<br />

(d) Detached sporangium showing basal<br />

mucilage ring. (e) Zygospore. (f,g) Stages in<br />

zygospore germination (e g) after Brefeld,<br />

1881).


EXAMPLES OF MUCORALES<br />

191<br />

Thamnidium<br />

The only species is T. elegans (Fig. 7.24), which<br />

grows in soil in cold and temperate regions, and<br />

on the dung of many different animals (Benny,<br />

1992). It is psychrophilic, continuing <strong>to</strong> grow<br />

at 1 2°C, with an optimum 18°C and a maximum<br />

at 27 31°C (Domsch et al., 1980). It has<br />

been reported from meat in cold s<strong>to</strong>rage.<br />

In culture, large terminal columellate sporangia<br />

are produced on tall sporangiophores which<br />

may also have repeatedly dicho<strong>to</strong>mous lateral<br />

branches bearing fewer-spored columellate or<br />

non-columellate sporangiola. The sporangiola<br />

may also be borne on separate branch systems.<br />

Low temperature and light induce the formation<br />

of sporangia as opposed <strong>to</strong> sporangiola. During<br />

the development of the sporangiophores, spiral<br />

growth occurs as in Phycomyces (see Fig. 7.6).<br />

Electron microscopy studies of the development<br />

of sporangia and sporangiola show that<br />

they develop in essentially the same way<br />

(J. Fletcher, 1973a,b). At maturity the columellate<br />

sporangia become converted in<strong>to</strong> sticky sporangial<br />

drops. In contrast, the sporangiola are easily<br />

detached in wind tunnel experiments. A change<br />

from damp <strong>to</strong> dry air leads <strong>to</strong> increased liberation<br />

of sporangiola (Ingold & Zoberi, 1963).<br />

Thamnidium elegans is heterothallic and forms<br />

zygospores resembling those of Mucor or Rhizopus,<br />

but they are produced best at low temperatures<br />

such as 6 7°C and not at 20°C (Hesseltine &<br />

Anderson, 1956).<br />

7.3.4 Chae<strong>to</strong>cladiaceae<br />

The family Chae<strong>to</strong>cladiaceae contains two<br />

genera, the facultatively mycoparasitic Chae<strong>to</strong>cladium<br />

and the saprotrophic Dicho<strong>to</strong>mocladium.<br />

Their fertile hyphae are branched and bear<br />

monosporous sporangiola on fertile vesicles.<br />

The main branches terminate in sterile spines<br />

(Benny & Benjamin, 1993). Whilst species of<br />

Chae<strong>to</strong>cladium are believed <strong>to</strong> be psychrophilic<br />

and are rarely collected within the tropics, all<br />

known species of Dicho<strong>to</strong>mocladium have been<br />

recorded only in tropical areas (Kirk, 1993).<br />

Chae<strong>to</strong>cladium<br />

In Chae<strong>to</strong>cladium (Fig. 7.25) there are no Mucor-like<br />

sporangia. Sporangiola, each containing a single<br />

spore, are borne on lateral branches which<br />

end in spines. Such monosporous sporangiola<br />

are sometimes termed conidia. There are two<br />

species, C. jonesii and C. brefeldii, both parasitic<br />

on other Mucorales (Benny & Benjamin, 1976),<br />

especially on Mucor or Pilaira growing on dung.<br />

At the point of attachment <strong>to</strong> the host there<br />

are numerous yellow galls. These are unique<br />

bladder-like outgrowths which contain nuclei<br />

of both the host and the parasite in a common<br />

cy<strong>to</strong>plasm. Chae<strong>to</strong>cladium does not form haus<strong>to</strong>ria<br />

and has been described as a fusion biotroph<br />

(Jeffries & Young, 1994). Both Chae<strong>to</strong>cladium spp.<br />

can, however, be cultured on standard agar<br />

media in the absence of a host. They are<br />

heterothallic. Chae<strong>to</strong>cladium brefeldii is heterogametangic,<br />

forming zygospores resembling<br />

Zygorhynchus. Burgeff (1920, 1924) has claimed<br />

that a given strain of Chae<strong>to</strong>cladium can only<br />

parasitize one of the two mating type strains<br />

of heterothallic Mucor spp., suggesting that the<br />

parasitic habit of fungi such as Chae<strong>to</strong>cladium<br />

may have originated from attempted copulation<br />

with other members of the Mucorales. Jeffries<br />

and Young (1994) believed that contact is truly<br />

mycoparasitic and not pseudosexual.<br />

7.3.5 Choanephoraceae<br />

This is probably the only current family in the<br />

Mucorales <strong>to</strong> be monophyletic (O’Donnell et al.,<br />

2001). Members of the Choanephoraceae are<br />

essentially tropical in their distribution. There<br />

are three genera of which the best-known are<br />

Blakeslea and Choanephora (Kirk, 1984). Asexual<br />

reproduction is by sporangia and sporangiola.<br />

The sporangia which have brown persistent walls<br />

are usually columellate and often hang downwards.<br />

They contain dark brown sporangiospores<br />

with a striate wall and bristle-like appendages at<br />

each end. The sporangiola contain one or a few<br />

spores, also with brown striate walls and with<br />

(Blakeslea) or without (Choanephora) polar appendages.<br />

The dark sporangium walls and the dark<br />

walls of the sporangiospores (due <strong>to</strong> melanin<br />

and carotenoid pigments), both unusual features<br />

in the Mucorales, may have evolved as a protection<br />

against the mutagenic and oxidizing UV<br />

light and may help <strong>to</strong> explain the tropical


192 ZYGOMYCOTA<br />

Fig 7.24 Thamnidium elegans.<br />

(a) Sporangiophore showing terminal<br />

sporangium and lateral branches bearing<br />

sporangiola. (b) Dehisced sporangium<br />

showing columella and spores. (c) Immature<br />

terminal sporangium showing the columella.<br />

(d) Base of sporangiophore with dicho<strong>to</strong>mous<br />

branches bearing sporangiola.<br />

(e) Sporangiola. Note the absence of<br />

a columella in these sporangiola. (b d) <strong>to</strong><br />

same scale.<br />

Fig 7.25 Chae<strong>to</strong>cladium brefeldii. (a) Habit<br />

sketch <strong>to</strong> show branches ending in spines<br />

and bearing lateral sporangiola.<br />

(b) Branch showing spine and sporangiola.<br />

(c) Hypha of Pilaira anomala bearing<br />

bladder-like outgrowths following<br />

parasitism by Chae<strong>to</strong>cladium.


EXAMPLES OF MUCORALES<br />

193<br />

distribution of these fungi (Kirk, 1993). The<br />

zygospores proper, when extruded from their<br />

zygosporangia, also have striate walls.<br />

Choanephora<br />

Choanephora cucurbitarum is a weak pathogen<br />

causing soft rot and wet rot diseases of a wide<br />

range of tropical and subtropical plants such as<br />

okra, chilli pepper, cowpea and Amaranthus. It<br />

also grows on decaying flowers of various kinds.<br />

Infection of male inflorescences of Ar<strong>to</strong>carpus<br />

integer (Moraceae) by Choanephora attracts gall<br />

midges which feed on the mycelium and build<br />

up large populations on the decaying flesh of the<br />

inflorescence. The gall midges are probably<br />

involved in pollination of the female inflorescences<br />

of Ar<strong>to</strong>carpus (Sakai et al., 2000).<br />

Asexual reproduction is by two types of<br />

structure, drooping multisporous sporangia and<br />

monosporous sporangiola (‘conidia’) borne on<br />

separate sporangiophores (Fig. 7.26). The development<br />

of sporangia is stimulated by growth on<br />

carbon-limited media and temperatures around<br />

30°C, whilst the optimum temperature for<br />

sporangiolum formation is around 25°C. Light<br />

is essential for sporulation. The sporangia are<br />

columellate or non-columellate and dehisce in<strong>to</strong><br />

two halves along a line of weakness. The<br />

sporangiospores have brown walls with longitudinal<br />

grooves appearing as striations, and<br />

bear a group of hyaline tapering appendages at<br />

each pole. These appendages may play a role in<br />

the dispersal of the spores in water films since<br />

they only become extended if the sporangium<br />

dehisces in water (Higham & Cole, 1982).<br />

Sporangiola develop on globose vesicles at the<br />

tips of separate sporangiophores, each of which<br />

may bear about 100 sporangiola. The sporangiolum<br />

is multinucleate and the spore within<br />

it develops a separate thick, brown, ridged<br />

Fig 7.26 Choanephora cucurbitarum.<br />

(a) Sporangiophore with drooping sporangium.<br />

(b) Sporangiophore (‘conidiophore’) with<br />

numerous monosporous sporangiola (‘conidia’).<br />

(c) Apex of conidiophore showing swollen<br />

vesicles bearing conidia. (d) Dehisced<br />

sporangium showing striate spores with<br />

terminal appendages. (e) Conidium.<br />

(f) Sporangiospore. (c) and (d) <strong>to</strong> same scale.


194 ZYGOMYCOTA<br />

(i.e. striate) wall inside the thin sporangiolum<br />

wall which clings <strong>to</strong> it and conforms <strong>to</strong> its shape,<br />

making it difficult <strong>to</strong> discern that the spore<br />

wall is distinct (Higham & Cole, 1982). The spore<br />

inside the sporangiolum has no appendages.<br />

Choanephora cucurbitarum is heterothallic. Its<br />

zygosporangia develop from intertwined zygophores<br />

and are held in place between <strong>to</strong>ngs-like<br />

suspensors (Kirk, 1977; Chang et al., 1984). The<br />

zygosporangial wall is thin and may flake off<br />

or fracture <strong>to</strong> reveal the striate wall of the<br />

enclosed zygospore.<br />

Blakeslea<br />

Blakeslea trispora, which has been isolated from<br />

cowpeas, <strong>to</strong>bacco and cucumber leaves, forms<br />

two kinds of asexual reproductive structure in<br />

culture: nodding columellate or non-columellate<br />

sporangia with brown, faintly striate spores<br />

which usually bear bristle-like appendages, and<br />

non-columellate sporangiola borne in large<br />

numbers on globose vesicles (Fig. 7.27). The<br />

sporangiola contain 2 5 (typically 3) distinctly<br />

striate, dark brown spores which also have<br />

bristle-like appendages. The production of the<br />

Fig 7.27 Blakeslea trispora.<br />

(a) Sporangiophores with globose<br />

terminal vesicles bearing<br />

sporangiola containing three or<br />

four spores. (b) A dehisced<br />

sporangiolum showing two spores<br />

released from it. Note the striate<br />

wall, the polar spore appendages<br />

and the splitting of the<br />

sporangiolum wall in<strong>to</strong> two halves.<br />

(c) Sporangiophore with a drooping<br />

sporangium. No columella was<br />

observed. (d) Dehisced sporangium<br />

also lacking a columella. Note the<br />

split sporangial wall and the<br />

sporangiospores with striate walls<br />

and polar appendages.


EXAMPLES OF MUCORALES<br />

195<br />

nodding sporangia is enhanced in culture by<br />

growth at a temperature of 30°C and that of<br />

sporangiola by a temperature of 26°C, with<br />

mixed sporulation at 28°C (Tereshina &<br />

Feofilova, 1995). The number of spores within<br />

sporangiola is affected by nutrition, and when<br />

grown on media with limiting nutrient content<br />

the sporangiola may contain only a single spore,<br />

thus resembling Choanephora. InB. unispora the<br />

sporangiola generally contain only one spore,<br />

rarely two. The sporangiola of B. trispora are<br />

readily detached by wind and break open in<br />

water like the two halves of a bivalve shell <strong>to</strong><br />

release the spores which are carried by insects<br />

from one plant <strong>to</strong> another (Fig. 7.27b). Blakeslea<br />

trispora is heterothallic and has brown striate<br />

zygospores resembling those of Choanephora<br />

(Mistry, 1977).<br />

The Choanephoraceae have been the subject<br />

of physiological investigations. An interesting<br />

phenomenon observed in intra- and inter-specific<br />

crosses is that the production of b-carotene is<br />

markedly enhanced when (þ) and ( ) strains are<br />

mated on liquid media, as compared with<br />

production from either strain grown singly.<br />

Commercial production of b-carotene and lycopene<br />

from fermentations of mixed cultures of<br />

(þ) and ( ) strains of B. trispora is possible (Mehta<br />

et al., 2003). The discovery that b-carotene<br />

production can be stimulated by an acid fraction<br />

of culture filtrates from mixed cultures of<br />

B. trispora led <strong>to</strong> the discovery of trisporic acid<br />

as the sex hormone of Mucorales (see p. 173).<br />

Syncephalastrum<br />

Syncephalastrum racemosum (Fig. 7.28) can be<br />

isolated from soil and dung in tropical and<br />

subtropical areas (Domsch et al., 1980). It grows<br />

rapidly in culture over a wide range of temperatures<br />

(7 40°C) and is mainly saprotrophic, but<br />

has been implicated in mucormycosis in human<br />

and animal hosts. It has also been isolated from<br />

foodstuff, cereal grains, other seeds and spices.<br />

In culture it forms aerial branches terminating<br />

in club-shaped or spherical vesicles. The vesicles<br />

are multinucleate and bud out all over their<br />

surface <strong>to</strong> form cylindrical outgrowths, the merosporangial<br />

primordia. In<strong>to</strong> these outgrowths<br />

one or perhaps several nuclei pass, and nuclear<br />

division continues. The cy<strong>to</strong>plasm in the merosporangium<br />

cleaves in<strong>to</strong> a single row of 5 10<br />

7.3.6 Syncephalastraceae<br />

A characteristic feature of this family is that<br />

asexual reproduction occurs by means of cylindrical<br />

sporangia containing typically a single row<br />

of sporangiospores. Such sporangia are termed<br />

merosporangia and are formed in groups on<br />

inflated vesicles (Benjamin, 1966). Merosporangia<br />

appear <strong>to</strong> have evolved independently in the<br />

Pip<strong>to</strong>cephalidaceae (Zoopagales) (see p. 201).<br />

There is only a single genus, Syncephalastrum, in<br />

the Syncephalastraceae. DNA sequence analysis<br />

indicates close relationships with certain genera<br />

traditionally classified in Mucoraceae and<br />

Thamnidiaceae (O’Donnell et al., 2001).<br />

Fig 7.28 Syncephalastrum racemosum. (a) Sporangiophore<br />

bearing a vesicle and numerous merosporangia.<br />

(b) Merosporangia and merospores.


196 ZYGOMYCOTA<br />

sporangiospores, each with 1 3 nuclei. The<br />

cleavage process is similar <strong>to</strong> that found in<br />

other Mucorales (Fletcher, 1972). The sporangial<br />

wall shrinks at maturity so that the spores<br />

appear in chains reminiscent of Aspergillus.<br />

Occasionally the merospores may lie in more<br />

than a single row. The spore heads remain dry<br />

and entire rows of spores (spore rods) are<br />

detached by wind (Ingold & Zoberi, 1963).<br />

Syncephalastrum racemosum is heterothallic and<br />

forms zygospores resembling those of other<br />

Mucorales.<br />

7.3.7 Cunninghamellaceae<br />

In this family, asexual reproduction is entirely by<br />

means of monosporous sporangiola. Sporangia<br />

are not formed. There is a single genus.<br />

Cunninghamella<br />

There are about 12 species of Cunninghamella,<br />

found in soil in the warmer regions of the world,<br />

e.g. the Mediterranean and subtropics (Domsch<br />

et al. 1980; Zheng & Chen, 2001). Cunninghamella<br />

elegans and C. echinulata are saprotrophs but C.<br />

bertholettiae is a serious, sometimes fatal, human<br />

pathogen. Cunninghamella echinulata may also be<br />

a destructive mycoparasite of Rhizopus arrhizus.<br />

Cunninghamella elegans and C. echinulata have been<br />

used in a wide range of biotransformations of<br />

pharmaceutical products (Kieslich, 1997). DNA<br />

sequence studies have grouped C. echinulata with<br />

some of the genera traditionally classified in<br />

Mucoraceae (O’Donnell et al., 2001), but comparisons<br />

of fatty acid and cell wall composition of<br />

Cunninghamella japonica and Blakeslea trispora<br />

have suggested that Cunninghamella is related <strong>to</strong><br />

members of the Choanephoraceae, a conclusion<br />

reached also on morphological criteria by some<br />

other workers.<br />

The sporangiola of Cunninghamella are hyaline<br />

and clustered on globose vesicles (ampoules)<br />

on branched or unbranched sporangiophores<br />

(Fig. 7.29). They are sometimes referred <strong>to</strong> as<br />

conidia, but details of their development indicate<br />

that they are best interpreted as one-spored<br />

sporangiola. Khan and Talbot (1975) have studied<br />

sporangiolum development in C. echinulata. The<br />

ampoules are club-shaped, globose or pearshaped<br />

and bear spherical sporangiola, each<br />

arising from a tubular denticle. Localized areas<br />

of weakness in the ampoule wall, yielding <strong>to</strong><br />

turgor pressure, blow out <strong>to</strong> form the denticles.<br />

The wall is two-layered in the ampoule and<br />

denticle, but single in the developing sporangiolum<br />

where it develops hollow spines all over the<br />

surface. Within the sporangiolum wall a twolayered<br />

wall develops around the multinucleate<br />

pro<strong>to</strong>plast. Hawker et al. (1970) have studied the<br />

structure and germination of sporangiola in<br />

a species of Cunninghamella. Here, <strong>to</strong>o, the wall<br />

Fig 7.29 Cunninghamella echinulata. (a) A<br />

simple and a branched sporangiophore.<br />

(b,c) Apices of sporangiophores showing<br />

expanded vesicles developing<br />

sporangiola. (d) Apex of mature<br />

sporangiophore with cluster of<br />

attached and two detached<br />

spiny-walled sporangiola. Scale bar (a) ¼<br />

25 mm, (b d) ¼ 10 mm.


EXAMPLES OF MUCORALES<br />

197<br />

of the ungerminated sporangiolum consisted of<br />

at least two layers, the outermost layer relatively<br />

thin, enclosing a distinct thicker inner layer. On<br />

germination only the inner layer extends as a<br />

germ tube.<br />

The zygospores of Cunninghamella resemble<br />

those of Mucor.<br />

7.3.8 Mortierellaceae<br />

The distinctive feature of this family is that the<br />

sporangiophore produces only a rudimentary<br />

columella or lacks it al<strong>to</strong>gether. In the most<br />

frequently encountered genus, Mortierella, zygospores<br />

are often heterogametangic and may be<br />

naked or enclosed in a weft of mycelium. The<br />

family, which includes about seven genera, has<br />

been monographed by Zycha et al. (1969). DNA<br />

sequence analyses indicate relationships <strong>to</strong><br />

certain genera usually placed in Mucoraceae<br />

(O’Donnell et al., 2001; Tanabe et al., 2004;<br />

see Fig. 7.1), although many authors regard<br />

Mortierella and its allies as a separate order,<br />

Mortierellales.<br />

Mortierella<br />

About 90 species of Mortierella are known mainly<br />

from soil, the rhizosphere, and plant or animal<br />

remains in contact with soil (Gams, 1977;<br />

Domsch et al., 1980). These fungi can be isolated<br />

readily on nutrient-poor media which prevent<br />

the growth of more vigorous moulds. Many<br />

species are psychrophilic and may comprise the<br />

bulk of fungal isolates from soil if the isolation<br />

media are incubated near 0°C (Carreiro & Koske,<br />

1992). Mortierella wolfii is associated with mycotic<br />

abortion in cattle and can be isolated from the<br />

placenta and foetal s<strong>to</strong>mach contents and from<br />

liver. In nature it grows in warm soils, overheated<br />

silage and rotten hay and can grow well<br />

at 40 42°C (Austwick, 1976; Domsch et al., 1980).<br />

Certain species of Mortierella, e.g. M. alpina, have<br />

been used in fermentations as catalysts of<br />

biotransformations in the production of pharmaceuticals<br />

(Kieslich, 1997). Another focus<br />

of biotechnological interest is their accumulation<br />

of lipid, notably polyunsaturated fatty<br />

acids (PUFAs) which are of nutritional value<br />

(Dyal & Narine, 2005). These are also produced<br />

by thraus<strong>to</strong>chytrids (see p. 73). The genus<br />

Mortierella is polyphyletic, and many of the<br />

best-known species, including M. isabellina,<br />

M. ramanniana and M. vinacea, are now placed<br />

in other genera such as Micromucor or Umbelopsis<br />

(Meyer & Gams, 2003).<br />

The mycelium of most species of Mortierella<br />

is fine and, in agar culture, often shows a<br />

characteristic series of fan-like zones. Cultures<br />

frequently have a garlic-like odour. The sporangia<br />

are borne on branched or unbranched tapering<br />

sporangiophores (Fig. 7.30a). The sporangium wall<br />

is delicate and may collapse around the spores.<br />

There is no protruding columella (Fig. 7.30b).<br />

Frequently the entire sporangium is detached. In<br />

a number of species, and also dependent upon<br />

environmental conditions, there may be only<br />

one or a few spores per sporangium (Figs. 7.30c,d).<br />

Asexual reproduction may also include the<br />

formation of sessile, intercalary chlamydospores<br />

which are not dispersed but remain in<br />

the soil when their subtending mycelium<br />

breaks down (Fig. 7.30e). Stylospores are also<br />

produced; unfortunately this term has been used<br />

for two different, non-homologous structures. In<br />

its original application by van Tieghem, it<br />

referred <strong>to</strong> aerial chlamydospores, i.e. relatively<br />

thick-walled, stalked spores as seen, for example,<br />

in M. polycephala (Domsch et al., 1980). In other<br />

species, classified in the section Stylospora, e.g. M.<br />

humilis and M. zonata (Gams, 1977), single-spored<br />

sporangiola (Fig. 7.31a) have been termed stylospores.<br />

On detachment of the sporangiolum, the<br />

remnants of the sporangiolum wall can often be<br />

seen at the tip of the sporangiophore (Domsch<br />

et al., 1980). In some species, e.g. M. stylospora and<br />

M. zonata, only sporangiola are present and true<br />

sporangia are lacking. Mortierella chlamydospora<br />

also lacks true sporangia, reproducing asexually<br />

by intercalary smooth or stalked echinulate<br />

chlamydospores and sexually by zygospores<br />

(Ansell & Young, 1982).<br />

The zygospores of Mortierella spp. may be<br />

naked or surrounded by a partial or complete<br />

investment of sterile hyphae (see Figs. 7.31b,c).<br />

Of the 90 species, 26 are known <strong>to</strong> form zygospores,<br />

half of which are homothallic (Watanabe<br />

et al., 2001). It is likely that the majority of the<br />

remaining species will prove <strong>to</strong> be heterothallic.


198 ZYGOMYCOTA<br />

Fig 7.30 Mortierella hyalina. (a) Branched sporangiophore as viewed with the dissection microscope. (b) Intercalary chlamydospore.<br />

(c) One-spored sporangium (sporangiolum) showing separate walls, the outer belonging <strong>to</strong> the sporangium and the inner <strong>to</strong> the<br />

sporangiospore. (d) Multi-spored sporangium with the sporangial wall disintegrating. (e) Apex of a sporangiophore in which the<br />

sporangium has dehisced, leaving fragments of the sporangial wall as a frill. Note the absence of a bulging columella. (b e) <strong>to</strong> same<br />

scale. Reprinted from Weber and Tribe (2003), with permission from Elsevier.<br />

Zygospore production takes place in culture,<br />

often embedded in the agar, on media with a<br />

relatively poor nutrient content which<br />

discourages profuse development of aerial mycelium.<br />

A common feature is that zygospores are<br />

heterogametangic, one suspensor being considerably<br />

larger than the other. The smaller progametangium<br />

or gametangium does not enlarge<br />

and may disappear soon after plasmogamy. The<br />

early development of such heterogametangic<br />

zygospores is illustrated in the heterothallic<br />

M. umbellata (Fig. 7.32; Degawa & Tokumasu,<br />

1998). In this species, hyphal coiling occurs at the<br />

point of contact of compatible mycelia, followed<br />

by the development of club-shaped progametangia<br />

which grow parallel and closely appressed <strong>to</strong><br />

each other. One, the macroprogametangium,<br />

soon becomes much larger than the other, the<br />

microprogametangium. In each progametangium<br />

a septum delimits a terminal gametangium<br />

from a suspensor. The macrogametangium<br />

and macrosuspensor both enlarge considerably,


EXAMPLES OF MUCORALES<br />

199<br />

Fig 7.31 (a) Mortierella zonata sporangiola, one germinating. (b,c) Mortierellarostafinskii (after Brefeld,1876). (b) Developing zygospore.<br />

(c) Older zygospore surrounded by a weft of hyphae. (d) Mortierella epigama zygospore with unequal suspensors arising<br />

from a common branch showing that this fungus is homothallic and heterogametangic.The zygospore proper, lying within the<br />

zygosporangium, has a thick undulating wall.<br />

appearing as two contiguous spheres (Figs.<br />

7.32e,f). Eventually the macrogametangium<br />

becomes a zygosporangium, its wall ornamented<br />

with small warts, and containing a smooth,<br />

thick-walled zygospore.<br />

The heterothallic M. indohi is also heterogametangic,<br />

one progametangium being blunter<br />

and more rounded than the other. A cross wall<br />

develops only in this larger progametangium <strong>to</strong><br />

cut off a gametangium which enlarges and


200 ZYGOMYCOTA<br />

Fig 7.32 Mortierella umbellata, stages in zygospore<br />

development (traced from Degawa & Tokumasu,1998).<br />

(a) Developing progametangia with<br />

micro-progametangium on the left. (b) Swelling of<br />

progametangia. (c) Septum formation in<br />

micro-progametangium, arrowed. (d) Septa have formed<br />

in both progametangia <strong>to</strong> delimit the terminal gametangia<br />

from sub-terminal suspensors. (e) The macrogametangium<br />

and macrosuspensor have enlarged. (f) Mature<br />

zygosporangium containing a thick-walled zygospore.<br />

becomes converted in<strong>to</strong> a subspherical zygosporangium<br />

with a dimpled wall. The narrower progametangium<br />

does not enlarge appreciably. It is<br />

not divided by a septum and remains as a lateral<br />

attachment <strong>to</strong> the zygosporangium (Ansell &<br />

Young, 1983, 1988). Delimitation of the zygosporangium<br />

by means of a single wall in only one of<br />

the fusing progametangia occurs in several other<br />

species of Mortierella (Kuhlman, 1972).<br />

Mortierella capitata shows an unusual mode of<br />

zygospore development (Degawa & Tokumasu,<br />

1997). It is heterothallic and heterogametangic,<br />

with two mating types designated A and B. When<br />

compatible vegetative hyphae meet in culture,<br />

their tips swell <strong>to</strong> form progametangia. The<br />

hyphal tips from strain B are always larger<br />

than those from A and are designated as macroprogametangia.<br />

The narrower hyphae from<br />

strain A (microprogametangia) coil around the<br />

macroprogametangia, branch dicho<strong>to</strong>mously<br />

and become septate, resulting in the formation<br />

of microsuspensors and microgametangia. A<br />

septum divides the terminal macrogametangium<br />

from its macrosuspensor. The macrogametangium<br />

becomes the zygosporangium and eventually<br />

contains a thick-walled hyaline zygospore.<br />

The macrosuspensor elongates and persists so<br />

that the mature zygosporangium appears at the<br />

end of a long stalk surrounded at its base by the<br />

coiled microsuspensors. Apart from the other<br />

unusual features of this developmental process,<br />

M. capitata is distinctive in that the morphology<br />

of its gametangia is linked <strong>to</strong> mating type, i.e.<br />

the formation of macrogametangia occurs only<br />

in the B strain and microgametangia in the A<br />

strain. This condition, termed morphological<br />

heterothallism, is comparatively rare in<br />

Zygomycota and in fungi generally.<br />

Complete investment of the zygospore by<br />

branching hyphae is a feature of M. rostafinskii<br />

(see Figs. 7.31b,c) and M. erice<strong>to</strong>rum (Kuhlman,<br />

1972). The zygospores proper in Mortierella are<br />

hyaline with thick smooth walls, sometimes<br />

showing coarse, undulating folds (see Fig.<br />

7.31d). Little is known about the germination of<br />

zygospores.<br />

7.4 Zoopagales<br />

The order Zoopagales contains soil- and dunginhabiting<br />

parasites of fungi and small terrestrial<br />

animals such as pro<strong>to</strong>zoa and nema<strong>to</strong>des.<br />

Reproduction is by conidia, merosporangia<br />

and zygospores. Benny et al. (2001) recognized


ZOOPAGALES<br />

201<br />

5 families and 20 genera but we shall study only<br />

the Pip<strong>to</strong>cephalidaceae, earlier classified in the<br />

Mucorales.<br />

7.4.1 Pip<strong>to</strong>cephalidaceae<br />

This family includes Pip<strong>to</strong>cephalis and Syncephalis,<br />

both mycoparasites. DNA sequence analysis<br />

suggest that these two genera are not closely<br />

related (Tanabe et al., 2000). Pip<strong>to</strong>cephalis is a<br />

biotrophic haus<strong>to</strong>rial parasite which needs the<br />

presence of a susceptible host for good growth<br />

and reproduction (Manocha, 1975), although on<br />

certain agar media Pip<strong>to</strong>cephalis spores will<br />

germinate and give rise <strong>to</strong> a limited mycelium<br />

producing dwarf sporangiophores. The spores so<br />

formed are unable <strong>to</strong> germinate if transferred <strong>to</strong><br />

fresh agar, but they do germinate and infect a<br />

suitable host fungus if one is present. Syncephalis<br />

develops intrahyphal hyphae within the host<br />

mycelium and can be grown more readily in<br />

culture if supplied with appropriate nutrients<br />

(Jeffries & Young, 1994).<br />

Pip<strong>to</strong>cephalis<br />

Most of the 20 or so known species of Pip<strong>to</strong>cephalis<br />

(Gr. pip<strong>to</strong> ¼ <strong>to</strong> fall, kephale ¼ head) parasitize the<br />

mycelium of Mucorales, with P. xenophila exceptional<br />

in its ability <strong>to</strong> infect members of the<br />

Ascomycota. Species of Pip<strong>to</strong>cephalis are most<br />

abundant in the surface layers of soils where<br />

there is a rapid recycling of organic matter,<br />

such as in woodland and in grazed grassland<br />

(Richardson & Leadbeater, 1972). They also parasitize<br />

Mucorales on dung. A characteristic habitat<br />

for P. freseniana is herbivore dung <strong>to</strong>wards the<br />

end of the fruiting phase of Mucor and Pilaira.<br />

From an infected host mycelium Pip<strong>to</strong>cephalis<br />

develops an erect dicho<strong>to</strong>mous sporangiophore<br />

(Fig. 7.33a). Swollen nodulose (knobbly) head<br />

cells form at the tips of the branches (see<br />

Fig. 7.33c), and from these cylindrical merosporangia<br />

radiate outwards. The merosporangia are<br />

thin-walled and usually contain from one <strong>to</strong><br />

several multinucleate merospores, arranged in<br />

a single row. Pip<strong>to</strong>cephalis unispora is unusual in<br />

that its merosporangia contain only a single<br />

sporangiospore. Its merosporangial wall encloses<br />

the sporangiospore which has a two-layered wall<br />

and may contain 1 3 nuclei (Jeffries & Young,<br />

1975). At maturity Pip<strong>to</strong>cephalis merosporangia<br />

behave in two different ways (Ingold & Zoberi,<br />

1963). In some species the thin sporangial wall<br />

collapses around the spores which remain<br />

attached <strong>to</strong>gether as spore rods, appearing as<br />

short chains (see Fig. 7.33c). Alternatively, as in<br />

P. freseniana, the merosporangial wall becomes<br />

diffluent and all the spores in a head collapse<br />

<strong>to</strong> form a spore drop. In some species the whole<br />

head cell with its attached merospores becomes<br />

detached at maturity. All types of propagule can<br />

be dispersed by wind.<br />

On germination sporangiospores swell and<br />

emit one <strong>to</strong> several germ tubes (McDaniell &<br />

Hindal, 1982). There is a chemotropic attraction<br />

Fig 7.33 Pip<strong>to</strong>cephalis virginiana. (a) Habit<br />

sketch <strong>to</strong> show dicho<strong>to</strong>mous sporangiophore.<br />

(b) Head cell and intact merosporangia.<br />

(c) Head cells showing breakdown of<br />

merosporangia <strong>to</strong> form chains of spores.<br />

(d) Spore germination and formation of<br />

appressorium on a host hypha.<br />

(e) Appressorium and branched<br />

haus<strong>to</strong>rium on host hypha.The parasite<br />

mycelium is branched and extending <strong>to</strong> other<br />

host hyphae. (f) Zygospore.The fungus is<br />

homothallic. (b e) <strong>to</strong> same scale.


202 ZYGOMYCOTA<br />

of germ tubes <strong>to</strong>wards host hyphae (Fig. 7.33d),<br />

with preferential growth <strong>to</strong>wards the hyphal<br />

tips. On agar the chemotropic stimulus can be<br />

detected over distances as great as 5 mm (Evans<br />

& Cooke, 1982). Fimbriae extending outwards for<br />

up <strong>to</strong> 25 mm from the cell walls of potential host<br />

fungi may play a role in directing the growth of<br />

Pip<strong>to</strong>cephalis germ tubes <strong>to</strong>wards the host hyphae<br />

(Rghei et al., 1992). At the point of contact an<br />

appressorium develops, but in some combinations<br />

the parasite hyphae may coil around those<br />

of their host and several appressoria form. In<br />

successful host parasite combinations, the host<br />

wall is penetrated beneath the appressorium by<br />

mechanical and possibly also enzymatic means.<br />

An infection peg penetrates the host wall.<br />

Enclosed by the plasmalemma of the host cell,<br />

the tip of the penetration peg expands <strong>to</strong> form a<br />

haus<strong>to</strong>rium which may branch inside the host<br />

hypha. The haus<strong>to</strong>ria of Pip<strong>to</strong>cephalis have close<br />

similarity <strong>to</strong> those of biotrophic haus<strong>to</strong>rial parasites<br />

of plants (Manocha & Lee, 1971; Jeffries &<br />

Young, 1976). Nutrients taken up by the haus<strong>to</strong>rium<br />

are translocated <strong>to</strong> the germinating spore<br />

and its germ tubes may then grow out <strong>to</strong> form<br />

a mycelium which extends over the host hypha,<br />

producing further haus<strong>to</strong>ria. The distinctive<br />

biochemical features of the Mucorales which<br />

are correlated with their ability <strong>to</strong> support the<br />

growth of these mycoparasites are that their<br />

walls contain chi<strong>to</strong>san and that their cy<strong>to</strong>plasm<br />

is rich in the polyunsaturated fatty acid<br />

g-linolenic acid which is essential for growth<br />

of the mycoparasite (Manocha, 1975, 1981;<br />

Manocha & Deven, 1975).<br />

Recognition between the mycoparasite and<br />

its hosts operates on at least two levels, the cell<br />

wall and the pro<strong>to</strong>plast surface (Manocha et al.,<br />

1990). There are qualitative and quantitative<br />

differences in the carbohydrates present at the<br />

hyphal surface of host and non-host species.<br />

Attachment is favoured by the presence of two<br />

distinctive glycoproteins in the wall of susceptible<br />

host hyphae. These two glycoproteins act as<br />

subunits of an agglutinin which may serve as<br />

recep<strong>to</strong>r <strong>to</strong> a complementary protein in the<br />

mycoparasite (Manocha et al., 1997).<br />

Pip<strong>to</strong>cephalis virginiana readily infects young<br />

but not old cultures of Choanephora cucurbitarum.<br />

This is correlated with the fact that the wall of<br />

young hyphae of C. cucurbitarum is single-layered<br />

whilst that of older hyphae is double-layered.<br />

Although appressoria and penetration pegs<br />

develop on older hyphae, penetration of the<br />

inner layer of the cell wall is rarely successful.<br />

The inner wall layer develops a papilla opposite<br />

the point of attempted penetration (Manocha,<br />

1981). Similar findings were made when<br />

P. virginiana failed <strong>to</strong> penetrate the resistant<br />

species P. articulosus (Manocha & Golesorkhi,<br />

1981). Where successful penetration of a susceptible<br />

host occurs, the mycoparasite P. virginiana<br />

can suppress wall synthesis by the host in the<br />

vicinity of infection points, so overcoming one<br />

of its defence reactions (Manocha & McCullough,<br />

1985; Manocha & Zhonghua, 1997).<br />

The effects of Pip<strong>to</strong>cephalis spp. on the growth<br />

of their hosts are very variable (Curtis et al., 1978).<br />

In some combinations the rate of growth of<br />

dual cultures was not significantly different<br />

from that of uninfected hosts, in others it was<br />

reduced, whilst in yet others it was enhanced.<br />

These effects are temperature-dependent. Growth<br />

and sporulation of the coprophilous fungus<br />

Pilaira anomala were reduced in culture when<br />

infected by P. fimbriata or P. freseniana (Wood &<br />

Cooke, 1986). A curious effect was found in<br />

culture when P. fimbriata challenged its normally<br />

susceptible host Mycotypha microspora. In the<br />

presence of P. fimbriata the host grew in a yeastlike<br />

state which was not infected. In contrast,<br />

the mycelial state of this fungus is readily<br />

infected (Evans et al., 1978).<br />

Most species of Pip<strong>to</strong>cephalis are homothallic<br />

(Leadbeater & Mercer, 1957). In culture zygospores<br />

are usually formed within the agar. The<br />

mature zygospore is a spherical dark brown<br />

sculptured globose cell held between two <strong>to</strong>ngshaped<br />

suspensors.<br />

7.5 En<strong>to</strong>mophthorales<br />

Many En<strong>to</strong>mophthorales are parasites of insects<br />

and other animals, whilst some parasitize<br />

desmids, nema<strong>to</strong>des or fern prothalli, or grow<br />

saprotrophically in plant litter, dung or soil.


ENTOMOPHTHORALES<br />

203<br />

An illustrated account of en<strong>to</strong>mopathogenic<br />

species has been provided by Samson et al.<br />

(1988) and a key <strong>to</strong> genera by Humber (1997).<br />

The major en<strong>to</strong>mopathogenic genera are Batkoa,<br />

Conidiobolus, En<strong>to</strong>mophaga, En<strong>to</strong>mophthora, Erynia,<br />

Furia, Massospora, Neozygites, Pandora and<br />

Zoophthora. Some of these insect pathogens hold<br />

promise for the control of insect pests, not least<br />

because many of them can be grown in culture,<br />

albeit on complex media containing ingredients<br />

such as sugars, egg yolk, yeast extract and milk<br />

(Wolf, 1981; Papierok & Hajek, 1997). The cells of<br />

En<strong>to</strong>mophthorales are uninucleate or coenocytic<br />

with chitinous walls, or they may exist in<br />

the bodies of insects as wall-less pro<strong>to</strong>plasts.<br />

The absence of a wall presumably reduces the<br />

elicitation of immune responses in their hosts<br />

(Dunphy & Nolan, 1982). Asexual reproduction in<br />

most genera is by means of forcibly discharged<br />

conidia, and on germination such conidia may<br />

develop a variety of secondary conidia. Sexual<br />

reproduction is by isogamous or anisogamous<br />

conjugation between uni- or multinucleate<br />

gametangia, <strong>to</strong> give a thick-walled zygospore.<br />

Azygospores may also be formed without conjugation,<br />

but it is likely that nuclear fusion and<br />

reduction division occur during their development<br />

(McCabe et al., 1984).<br />

Fossil evidence indicates that, as insect<br />

pathogens, members of the group were extant<br />

at least 25 million years ago. A well-preserved<br />

specimen of a winged termite probably infected<br />

with a species of En<strong>to</strong>mophthora has been found<br />

embedded in amber dated around the<br />

Oligocene Miocene border in the Dominican<br />

Republic (Poinar & Thomas, 1982).<br />

According <strong>to</strong> Benny et al. (2001), the order<br />

En<strong>to</strong>mophthorales consists of six families including<br />

the Basidiobolaceae. If this family is<br />

excluded, the remaining En<strong>to</strong>mophthorales<br />

appear <strong>to</strong> be monophyletic by DNA-based analysis<br />

(Jensen et al., 1998). In many phylogenetic<br />

schemes, Basidiobolus ranarum seems <strong>to</strong> be more<br />

closely related <strong>to</strong> Chytridiales and Neomastigales<br />

than <strong>to</strong> En<strong>to</strong>mophthorales (see Figs. 1.26, 7.1),<br />

and Cavalier-Smith (1998) has placed it in a<br />

separate order, the Basidiobolales. In the current<br />

context, we sacrifice these taxonomic details in<br />

favour of a better understanding of the<br />

Zygomycota as a whole, and therefore retain<br />

Basidiobolus in the En<strong>to</strong>mophthorales.<br />

Important criteria in the classification of<br />

En<strong>to</strong>mophthorales are the branched or unbranched<br />

nature of the conidiophores, whether<br />

the conidia are uninucleate or multinucleate,<br />

whether the wall of the conidium is single<br />

(unitunicate) or separates in<strong>to</strong> two layers<br />

(bitunicate), and the presence or absence of<br />

secondary conidia and their morphology<br />

(Humber, 1989).<br />

7.5.1 Basidiobolaceae: Basidiobolus<br />

Basidiobolus is the only genus in the<br />

Basidiobolaceae. The best-known species is B.<br />

ranarum, which has a worldwide distribution. It<br />

fruits on the dung of frogs, <strong>to</strong>ads, lizards, some<br />

insectivorous fish and mammals such as bats. It<br />

has also been found on the dung of kangaroos<br />

and wallabies (Speare & Thomas, 1985). If a frog<br />

is captured and placed in a jar with a little water,<br />

it will defaecate in due course and its dung can<br />

be filtered off. If the damp filter paper is placed<br />

in the lid of an inverted Petri dish containing a<br />

suitable agar medium (e.g. 1% pep<strong>to</strong>ne agar,<br />

pota<strong>to</strong>-dextrose agar, or cornmeal agar), conidia<br />

of B. ranarum will be shot upwards from the dung<br />

on<strong>to</strong> the agar surface, and within a few days<br />

coarsely septate colonies will become visible on<br />

the agar (Weber & Webster, 1998a). The presence<br />

of Basidiobolus and other ballis<strong>to</strong>sporic fungi such<br />

as Conidiobolus in surface soil and litter can<br />

also be disclosed by the ‘canopy’ technique.<br />

A suspension of soil is filtered and the filter<br />

paper, bearing a thin layer of soil, is placed in the<br />

lid of a Petri dish facing downwards over a<br />

suitable agar medium. The dish is illuminated<br />

from below and this encourages the discharge of<br />

conidia on<strong>to</strong> the agar (Smith & Callaghan, 1987;<br />

Callaghan, 2004).<br />

In agar cultures of Basidiobolus, the cy<strong>to</strong>plasm<br />

in the mycelium moves <strong>to</strong>wards the hyphal apex<br />

so that only a few terminal segments contain<br />

cy<strong>to</strong>plasm and a single large, prominent nucleus<br />

whilst the older segments are empty, being<br />

isolated by retraction septa (Fig. 7.34d). The<br />

cy<strong>to</strong>plasm-filled mycelial segments are termed<br />

hyphal bodies. Branching occurs immediately


204 ZYGOMYCOTA<br />

Fig 7.34 Basidiobolus ranarum. (a) Gut-stage cell from<br />

fresh frog dung. (b) Germination of gut-stage cells<br />

producing a coarse septate mycelium. (c) Hyphal apex<br />

with the terminal cells full of cy<strong>to</strong>plasm.The prominent<br />

nucleus in the apical cell is arrowed. (d) Apex of an older<br />

hypha in which only the two terminal cells contain<br />

cy<strong>to</strong>plasm; those behind are empty. (e) Branches arising<br />

beneath the septa of three subterminal cells. Scale bar:<br />

(a,b) ¼ 25 mm, (c) ¼ 50 mm, (d,e) ¼ 100 mm.<br />

behind the septum delimiting the apical<br />

segment, after mi<strong>to</strong>tic division of the nucleus<br />

(Fig. 7.34e). The conidiophores, which develop in<br />

a few days, are pho<strong>to</strong>tropic and resemble the<br />

sporangiophores of Pilobolus but bear a colourless<br />

pear-shaped <strong>to</strong> globose ballis<strong>to</strong>sporic conidium<br />

(Fig. 7.35a). O’Donnell (1979) interpreted the<br />

conidium as a monosporous sporangiolum but<br />

since it can, under certain conditions, cleave <strong>to</strong><br />

form endospores it may also be regarded as a<br />

modified sporangium. The conidium is uninucleate.<br />

A conical columella projects in<strong>to</strong> it. Beneath<br />

is a swollen sub-conidial vesicle containing liquid<br />

under turgor pressure. This is probably generated<br />

by a single large vacuole which fills most of<br />

the sub-conidial vesicle at maturity. A line of<br />

weakness can be detected as a slight constriction<br />

around the base of the vesicle, and when this<br />

ruptures the conidium and vesicle fly forward for<br />

a distance of 1 2 cm. The elastic upper portion<br />

of the vesicle contracts and the vacuolar sap<br />

within it squirts out backwards, so that it<br />

behaves as a minute rocket (Ingold, 1971).<br />

During their flight the conidium and the rocket<br />

mo<strong>to</strong>r (i.e. the vesicle) may be separated or the<br />

two parts remain attached <strong>to</strong> each other until<br />

landing (Figs. 7.35c,d).<br />

Conidium germination in B. ranarum<br />

Primary ballis<strong>to</strong>sporic conidia can germinate in a<br />

number of different ways depending on external<br />

conditions (Zahari & Ship<strong>to</strong>n, 1988; Waters &<br />

Callaghan, 1999).<br />

(1) By direct germination, producing one <strong>to</strong><br />

several germ tubes from which the vegetative<br />

mycelium develops (Fig. 7.35e). Germination of<br />

this type requires a nutrient concentration above<br />

that of 0.1% malt extract agar.


ENTOMOPHTHORALES<br />

205<br />

(2) Germination by repetition <strong>to</strong> form a secondary<br />

conidiophore with a ballis<strong>to</strong>sporic conidium.<br />

This is essentially similar <strong>to</strong> the primary<br />

conidium (Fig. 7.35a) and is produced under<br />

conditions of high water availability and low<br />

nutrient concentration. Secondary conidia may<br />

germinate by further repetition or in other ways.<br />

(3) Discharged ballis<strong>to</strong>sporic conidia formed<br />

in culture on certain media or located within<br />

the gut of the frog may cleave <strong>to</strong> form many<br />

endospores (sporangiospores, sometimes termed<br />

meris<strong>to</strong>spores), and these are released by dissolution<br />

of the original conidial wall (Fig. 7.36a;<br />

Dykstra, 1994).<br />

(4) Germination under somewhat drier conditions<br />

with a water activity at or below 0.995<br />

stimulates the development of capilliconidia<br />

or capillispores (Fig. 7.36c). The body of the<br />

capilliconidium may cleave by transverse and<br />

longitudinal septa <strong>to</strong> form endogenous segments<br />

(endospores, meris<strong>to</strong>spores) which are released<br />

by breakdown of the wall of the capilliconidium<br />

(Fig. 7.36d; Drechsler, 1956).<br />

Capilliconidia are so called because they are<br />

formed on long (over 0.3 mm), slender conidiophores.<br />

The conidia themselves are spindleshaped<br />

and apically beaked with a terminal<br />

globose adhesive droplet or hap<strong>to</strong>r. The material<br />

making up the hap<strong>to</strong>r is extruded through a<br />

narrow channel within the beak of the conidium.<br />

The droplet has unusual properties because it is<br />

not affected by water but rapidly spreads out <strong>to</strong><br />

form a film when in contact with a solid surface<br />

(Dykstra & Bradley-Kerr, 1994). The capilliconidia<br />

are easily detached from their conidiophores<br />

and may be dispersed by mites (Blackwell &<br />

Fig 7.35 Basidiobolus ranarum. (a) Conidiophore from<br />

culture. Note the conical columella and the swollen<br />

vesicle with a line of weakness around its base.<br />

(b) Primary conidium germinating <strong>to</strong> produce a<br />

secondary conidiophore and ballis<strong>to</strong>sporic conidium.<br />

(c) Discharged conidium with remnant of the vesicle<br />

attached. (d) Discharged conidium separated from the<br />

remnant of the vesicle. (e) Conidium germinating directly<br />

<strong>to</strong> form a septate mycelium.


206 ZYGOMYCOTA<br />

Fig 7.36 Basidiobolusranarum. (a) Stages in the<br />

development of endospores by primary<br />

ballis<strong>to</strong>conidia. Above: cy<strong>to</strong>plasmic contents cleaving<br />

<strong>to</strong> form two pro<strong>to</strong>plasts.Centre: sporangium<br />

containing over a dozen sporangiospores. Below:<br />

sporangium showing breakdown of sporangium wall<br />

(traced from Dykstra,1994). (b) Successive cleavage<br />

of pro<strong>to</strong>plasts from a primary ballis<strong>to</strong>sporic conidium<br />

placed on a rich agar medium <strong>to</strong> form the‘Palmella’<br />

stage. (c) Germination of a primary ballis<strong>to</strong>conidium<br />

<strong>to</strong> form a uninucleate secondary capilliconidium with<br />

a terminal beak which has extruded a sticky hap<strong>to</strong>r.<br />

(d) Capilliconidium which has divided <strong>to</strong> produce<br />

several endospores, some of which have been<br />

released following breakdown of its wall.One of the<br />

endospores is germinating. Scale bars: (a,b) ¼ 20 mm,<br />

(c,d) ¼ 12.5 mm.<br />

Malloch, 1989). Mites are ingested by beetles, the<br />

main vec<strong>to</strong>rs of B. ranarum, but other insects,<br />

spiders, millipedes, woodlice, worms and snails<br />

may also acquire conidia. These are ingested by<br />

vertebrate insectivores. Within the vertebrate<br />

gut, endospores are released from ballis<strong>to</strong>sporic<br />

conidia and capilliconidia.<br />

Endospores germinate in the vertebrate gut<br />

<strong>to</strong> form spherical, large, uninucleate, hyaline<br />

cells up <strong>to</strong> 20 mm. These were called the ‘Darm-<br />

Form’ (gut-stage) of B. ranarum by Levisohn (1927)<br />

who showed that a single ingested primary<br />

conidium can give rise <strong>to</strong> 50 60 division<br />

products (meris<strong>to</strong>spores) forming gut-stage cells<br />

(Fig. 7.34a). The concentration of Basidiobolus<br />

propagules builds up in the guts of the vertebrate<br />

vec<strong>to</strong>rs, and the fungus can be isolated<br />

from faeces of lizards up <strong>to</strong> 18 days after the<br />

animals are deprived of infected prey (Coremans-<br />

Pelseneer, 1973; Okafor et al., 1984). The gut-stage<br />

cells are voided with the faeces and can survive<br />

for several months under dry conditions. Under<br />

moist warm conditions they germinate <strong>to</strong> form a<br />

mycelium from which ballis<strong>to</strong>sporic conidiophores<br />

develop, whereas on certain media they<br />

enlarge and their contents undergo successive<br />

binary fission <strong>to</strong> form globose thick-walled cells.<br />

This state is sometimes termed the ‘Palmella’<br />

state (Fig. 7.36b) because of its superficial<br />

resemblance of a genus of green algae.


ENTOMOPHTHORALES<br />

207<br />

Basidiobolus microsporus, which grows in<br />

deserts in California, has a method of asexual<br />

reproduction not found in B. ranarum. Primary<br />

conidia can germinate directly or by repetition<br />

as in B. ranarum, but capilliconidia have not<br />

been found. However, under relatively dry<br />

conditions primary conidia may produce large<br />

numbers of exogenous obclavate spores (microspores)<br />

each attached by a separate pedicel <strong>to</strong><br />

the wall of the primary conidium. They have<br />

been interpreted as modified sporangiospores<br />

(Benjamin, 1962).<br />

In culture it has been found that light,<br />

especially blue light of wavelength 440 480 nm,<br />

stimulates conidial development and discharge<br />

in B. ranarum. The effect of light is <strong>to</strong> stimulate<br />

aerial growth from hyphal bodies within the<br />

medium, and the aerial hyphae which develop<br />

in the light become modified as conidiophores<br />

(Callaghan, 1969a,b).<br />

Sexual reproduction in B. ranarum<br />

Zygospores are formed following conjugation.<br />

The fungus is homothallic and development can<br />

be seen on certain agar media (e.g. Czapek-Dox<br />

agar) within 4 5 days in cultures derived from a<br />

single conidium. Zygospore development appears<br />

<strong>to</strong> occur most readily in the dark, and under these<br />

conditions the hyphal bodies become bicellular<br />

prior <strong>to</strong> developing in<strong>to</strong> zygospores (Callaghan,<br />

1969b). On either side of a septum, beak-like<br />

projections develop, and the single nucleus<br />

within each hyphal segment migrates in<strong>to</strong> the<br />

tip and divides there. One daughter nucleus is cut<br />

off by a septum in the terminal cell of the beak<br />

and later disintegrates, whereas the second<br />

nucleus migrates back in<strong>to</strong> the parent cell.<br />

Following this, one of the parent cells enlarges<br />

<strong>to</strong> several times the volume of the adjacent cell<br />

and a pore is formed connecting the two cells<br />

through the original septum separating them.<br />

A nucleus from the smaller cell passes through<br />

the pore and lies close <strong>to</strong> the nucleus of the larger<br />

cell. Nuclear fusion may occur directly or after<br />

a further division. The enlarged parent cell forms<br />

the zygospore which has a thick wall when<br />

mature (Fig. 7.38). Meiosis occurs within the<br />

mature zygospore <strong>to</strong> give four haploid nuclei, of<br />

which three usually degenerate. The mature<br />

zygospores of some isolates of B. ranarum have<br />

thick undulating walls of variable thickness,<br />

but in others the wall may be smooth. On germination<br />

the zygospore forms a germ tube<br />

or a conidiophore terminated by a ballis<strong>to</strong>sporic<br />

conidium. Capilliconidia may also develop from<br />

germinating zygospores (Dykstra & Bradley-Kerr,<br />

1994). The complicated and unusual life cycle<br />

of B. ranarum is illustrated in Fig. 7.37.<br />

Basidiobolus ranarum is an atypical zygomycete<br />

in that its mycelium becomes divided in<strong>to</strong> uninucleate<br />

segments. The nucleus is also unusually<br />

large, up <strong>to</strong> 25 mm, and this fact has led <strong>to</strong> several<br />

investigations of its cy<strong>to</strong>logy (e.g. Robinow, 1963;<br />

Tanaka, 1970; Sun & Bowen, 1972). The number<br />

of chromosomes has been estimated <strong>to</strong> be as high<br />

as 900, and the nucleus may be polyploid.<br />

Pathogenicity of B. ranarum<br />

Basidiobolus is probably not harmful <strong>to</strong> most<br />

insects and mites, although it has been isolated<br />

as a mass infection of mosqui<strong>to</strong>es, from termites<br />

and from larvae of Galleria (Krejzová, 1978). It was<br />

earlier thought <strong>to</strong> be harmless <strong>to</strong> reptiles and<br />

amphibians and there is no evidence of intestinal<br />

lesions in them. However, an epizootic cutaneous<br />

infection caused by B. ranarum has been reported<br />

from the dwarf African clawed frog, Hymenochirus<br />

curtipes (Groff et al., 1991). There are many reports<br />

of the isolation of B. ranarum from man and<br />

domestic animals such as horses (see Gugnani,<br />

1999; Ribes et al., 2000). Although several specific<br />

names have been applied <strong>to</strong> isolates pathogenic<br />

<strong>to</strong> humans and other mammals, the consensus is<br />

that they should be regarded as synonyms<br />

of B. ranarum (McGinnis, 1980). This view is<br />

supported by ribosomal DNA analysis (Nelson<br />

et al., 1990). Isolates from humans, unsurprisingly,<br />

are capable of growing at 37°C (Cochrane<br />

et al., 1989). Human disease caused by B. ranarum<br />

is more common in tropical and subtropical<br />

regions than in temperate zones. Infection is<br />

associated with subcutaneous swellings of<br />

affected areas of the lower limbs but rare intestinal<br />

infections are also known. It is assumed that<br />

the inoculum is usually soil-borne, and the use<br />

of fallen leaves in place of <strong>to</strong>ilet paper has<br />

sometimes been implicated as the cause of


208 ZYGOMYCOTA<br />

Fig 7.37 The eventful life cycle of Basidiobolus ranarum, not <strong>to</strong> scale. A beetle with attached or ingested ballis<strong>to</strong>conidia and<br />

capilliconidia is eaten by a frog. In the gut of the frog, both conidial types can undergo cleavage <strong>to</strong> form endospores, which germinate<br />

by enlargement <strong>to</strong> form the gut stage. After defaecation, gut-stage cells germinate <strong>to</strong> produce ballis<strong>to</strong>conidia or cleave <strong>to</strong> give the<br />

Palmella stage. Discharged ballis<strong>to</strong>conidia may germinate by repetition, by forming capilliconidia, or by emitting a hypha. Zygospore<br />

formation is initiatedby conjugation between two adjacent hyphal cells. Small open circles represent haploid nuclei; diploid nuclei are<br />

larger and split. Basidiobolusranarum is homothallic. Key events in the life cycle are plasmogamy (P), karyogamy (K) and meiosis (M).<br />

infections. In horses, infection of the nasal<br />

mucosa is again most probably from soil.<br />

7.5.2 Ancylistaceae: Conidiobolus<br />

There are about 30 species of Conidiobolus, the<br />

‘conidium thrower’. King (1977) has given keys<br />

and descriptions. Most of them grow saprotrophically<br />

in soil and litter and can be readily<br />

isolated by the canopy technique described for<br />

Basidiobolus (see p. 203) because they forcefully<br />

project their conidia. Several species have been<br />

isolated from basidiocarps of the Jew’s Ear<br />

fungus, Auricularia auricula-judae. Some are pathogenic<br />

<strong>to</strong> insects such as aphids and termites, and<br />

certain species cause disease in mammals<br />

including man. The characteristic feature of<br />

Conidiobolus is the formation of globose multinucleate<br />

primary conidia surrounded by a twolayered<br />

wall whose layers do not separate (Latgé<br />

et al., 1989). The primary conidia are projected<br />

by an eversion mechanism in which the inner<br />

wall of the double-walled conidiophore apex<br />

(‘columella’) suddenly rounds off because of<br />

different mechanical properties of the two wall<br />

layers. The force responsible is turgor pressure.<br />

After discharge, the tip of the columella can be<br />

seen projecting outwards as a conical papilla,<br />

and the base of the discharged conidium ends<br />

in a similar projection (Fig. 7.39b). Spores may<br />

be shot away for up <strong>to</strong> 4 cm. Germination of<br />

primary conidia may be by repetition producing<br />

the same spore type, by germ tubes (direct


ENTOMOPHTHORALES<br />

209<br />

Fig 7.38 Basidiobolus ranarum. Successive<br />

stages in the formation of zygospores.<br />

(a) Progametangia. (b) Young zygospore.<br />

(c) Mature zygospore.<br />

germination), by the formation, in some species,<br />

of numerous microconidia, or by the development<br />

of capilliconidia resembling those of<br />

Basidiobolus. Zygospores or azygospores have<br />

been reported and all species which reproduce<br />

sexually in this way are homothallic. It is<br />

believed that the cy<strong>to</strong>logical condition of the<br />

nuclei is haploid, as it is in other Zygomycota,<br />

and that karyogamy and meiosis are involved in<br />

the formation of zygospores and azygospores<br />

(McCabe et al., 1984).<br />

The best-known species of Conidiobolus is the<br />

cosmopolitan C. coronatus, a fungus which has<br />

been referred <strong>to</strong> under various names, e.g.<br />

En<strong>to</strong>mophthora coronata, Delacroixia coronata and<br />

Conidiobolus villosus. It grows readily and rapidly<br />

in agar culture, forming a septate mycelium and<br />

numerous pho<strong>to</strong>tropic conidiophores (Fig. 7.39)<br />

which shoot off conidia on<strong>to</strong> the lid of the Petri<br />

dish. Conidial discharge takes place both in the<br />

light and in the dark, but is enhanced by light<br />

(Callaghan, 1969a).<br />

The behaviour of a conidium on germination<br />

depends on pH, humidity, availability of light,<br />

and nutrients. If the conidium falls on a medium<br />

containing nutrients, it germinates by means of a<br />

germ tube, but on nutrient-poor media, such as<br />

water agar, it may develop in<strong>to</strong> a secondary<br />

conidiophore, forming a slightly smaller conidium<br />

(Fig. 7.39c). The secondary conidiophore<br />

develops from the illuminated side of a primary<br />

conidium, and the conidiophore which develops<br />

is pho<strong>to</strong>tropically orientated, but not very<br />

precisely (Page & Humber, 1973). Under conditions<br />

of reduced humidity the primary conidia<br />

may develop a cluster of globose microconidia<br />

(Fig. 7.39f). The entire cy<strong>to</strong>plasm of the primary<br />

conidium is evacuated by the expansion of a large<br />

vacuole in<strong>to</strong> numerous buds formed by localized<br />

softening of the primary conidium wall. At


210 ZYGOMYCOTA<br />

Fig 7.39 Conidiobolus<br />

coronatus. (a) Conidiophore<br />

with attached primary<br />

conidium. Note the columella<br />

which protrudes in<strong>to</strong> the body<br />

of the conidium. (b) Apex of<br />

conidiophore and conidium<br />

after discharge by eversion of<br />

the columella from inside the<br />

spore. (c) Primary conidium<br />

germinating by repetition <strong>to</strong><br />

produce a secondary conidium<br />

of similar type. (d) Conidium<br />

germinating directly, forming<br />

several germ tubes.<br />

(e) Conidium germinating <strong>to</strong><br />

produce a villose secondary<br />

conidium. (f) Conidia<br />

germinating <strong>to</strong> produce<br />

numerous secondary<br />

microconidia which are<br />

discharged by columellar<br />

eversion.<br />

maturity each microconidium is supported on a<br />

two-ply columella and projected by the eversion<br />

mechanism. In older cultures, primary conidia<br />

may form short conidiophores terminating in<br />

pear-shaped spiny-walled (villose) conidia, which<br />

are also projected by columellar eversion. The<br />

precise conditions under which spiny conidia are<br />

formed are not known. They have been interpreted<br />

as resting spores, and they germinate <strong>to</strong><br />

form a coarse septate mycelium. The possession of<br />

both microconidia and villose conidia is a<br />

combination unique <strong>to</strong> C. coronatus. Another<br />

type of resting spore may develop from ungerminated<br />

primary conidia which swell and form 2 3


ENTOMOPHTHORALES<br />

211<br />

layered thickened walls. Such conidia have been<br />

termed loriconidia (Gindin & Ben-Ze’ev, 1994).<br />

No zygospores have been reported for C. coronatus,<br />

structures resembling zygospores being interpreted<br />

as aerial chlamydospores.<br />

Conidiobolus coronatus is a parasite of aphids,<br />

termites and whiteflies attacking <strong>to</strong>bacco, cot<strong>to</strong>n<br />

and sweet pota<strong>to</strong>, as well as of waxmoths and<br />

some other insects (Gindin & Ben-Ze’ev, 1994;<br />

Bogus & Szczepanik, 2000). It should be regarded<br />

as a relatively primitive opportunistic pathogen.<br />

Infection of termites can occur by penetration of<br />

germ tubes through the exoskele<strong>to</strong>n, or via the<br />

oesophagus after ingestion of germinated conidia<br />

(Yendol & Paschke, 1965). Following infection<br />

of insects, death can occur within 2 days, probably<br />

by the production of <strong>to</strong>xins (Evans, 1989).<br />

A highly insecticidal 30 kDa protein has been<br />

found in mycelium and culture filtrates of<br />

C. coronatus (Bogus & Scheller, 2002). This and<br />

possibly other <strong>to</strong>xins induce damage <strong>to</strong> blood<br />

cells or early death in several insects when<br />

injected in<strong>to</strong> the haemocoel. In artificially<br />

infected waxmoths (Galleria mellonella), infection<br />

is followed by melanization of the host cuticle<br />

and damage <strong>to</strong> the Malpighian tubules with<br />

no evidence of tissue penetration (Bogus &<br />

Szczepanik, 2000). Conidiobolus coronatus and<br />

C. obscurus are being investigated as potential<br />

agents of biological control of insect pests.<br />

Conidiobolus coronatus is also pathogenic <strong>to</strong><br />

mammals such as horses, llama, chimpanzee<br />

and man. Human infections are most common in<br />

the moist tropics and subtropics, especially in<br />

male outdoor workers from the rain forests of<br />

West Africa. Although the mode of transmission<br />

has not been established it is probably by inhalation<br />

of spores which germinate in the nasal<br />

mucosa. Other species of Conidiobolus known <strong>to</strong><br />

infect vertebrates are C. incongruus and C.<br />

lamprauges. Isolates pathogenic <strong>to</strong> vertebrates<br />

grow readily at 37°C (Gugnani, 1992; Ribes<br />

et al., 2000).<br />

7.5.3 En<strong>to</strong>mophthoraceae<br />

Benny et al. (2001) included 12 en<strong>to</strong>mopathogenic<br />

genera in the family En<strong>to</strong>mophthoraceae<br />

(Gr. ‘insect destroyer’), of which we shall study<br />

the three most important, i.e. Erynia,<br />

En<strong>to</strong>mophthora and Furia.<br />

Erynia<br />

There are about 12 species of Erynia parasitic<br />

on terrestrial insects such as aphids and<br />

Lepidoptera, but some attack the aquatic larval<br />

stages of Diptera such as Simulium spp. (river<br />

blackflies), s<strong>to</strong>ne flies and caddis flies.<br />

Characteristic features of the genus are branched<br />

conidiophores bearing uninucleate, bitunicate<br />

primary conidia which are discharged by septal<br />

eversion. Germination of primary conidia is by<br />

the production of various types of secondary<br />

conidia. Tertiary conidia may also develop.<br />

Resting bodies (azygospores and zygospores)<br />

occur in some, but not all species. Attempts are<br />

being made <strong>to</strong> use Erynia neoaphidis <strong>to</strong> control<br />

aphid populations in field crops (Pell et al., 2001).<br />

Erynia neoaphidis<br />

This species, synonymous with En<strong>to</strong>mophthora<br />

aphidis, Pandora neoaphidis and Zoophthora neoaphidis,<br />

is the most widespread aphid pathogen of<br />

temperate regions and has been found on over<br />

70 species of aphids on annual and perennial<br />

crops. It also attacks aphids on non-cultivated<br />

plants, a common example being the nettle<br />

aphid, Microlophium (Fig. 7.40). Infected aphids<br />

are cream <strong>to</strong> brown in colour. They are attached<br />

on the ventral side <strong>to</strong> their plant host by fungal<br />

rhizoids and their bodies are distended. Within<br />

the body of an infected aphid there are numerous<br />

closely packed, wide, septate hyphal bodies<br />

(Fig. 7.41a). Widely spaced, thick-walled, long,<br />

awl-shaped pseudocystidia (Fig. 7.41b) pierce<br />

the cuticle and, surrounding them, numerous<br />

tightly packed, branched conidiophores emerge,<br />

usually made up of uninucleate segments<br />

(Figs. 7.41b,d; Brobyn & Wilding, 1977). The tip<br />

of the conidiophore is cut off by a two-ply<br />

septum <strong>to</strong> form a uninucleate primary conidium<br />

with a two-layered wall (Figs. 7.41d,e).<br />

Under humid conditions (relative humidity<br />

495%), primary conidia are discharged by<br />

septal eversion for a distance of about 1 cm,<br />

and detached conidia show a bulging papilla<br />

at their base. Violent discharge projects the<br />

conidia through the boundary layer of still air


212 ZYGOMYCOTA<br />

Fig 7.4 0 Carcass of a nettle aphid (Microlophium) incubated<br />

on tap water agar for12 h. Note the halo of discharged conidia<br />

of Erynia neoaphidis.<br />

surrounding the host so that they come under<br />

the influence of wind and gravity, falling at a<br />

velocity of about 1 cm s 1 (Hemmati et al., 2002).<br />

About 200 000 conidia are produced per cadaver<br />

of adult pea aphid over a period of 2 3 days.<br />

Immediately after discharge, primary conidia<br />

may germinate <strong>to</strong> produce secondary conidia<br />

which are wider and more ovate than the<br />

primary conidia (see Figs. 7.41e,f). Direct germination<br />

by the production of a germ tube from<br />

one or both ends of primary and secondary<br />

conidia also occurs (Fig. 7.41c). As in many<br />

En<strong>to</strong>mophthorales, cy<strong>to</strong>plasm is concentrated<br />

in<strong>to</strong> a few terminal cells, leaving empty intercalary<br />

segments cut off by retraction septa<br />

(Fig. 7.41c).<br />

Brobyn and Wilding (1977) and Butt et al.<br />

(1990) have described the process of infection of<br />

the pea aphid Acyrthosiphon pisum. Conidia adhere<br />

<strong>to</strong> any point on the aphid cuticle, often in<br />

clumps, and may germinate by forming secondary<br />

conidia or germ tubes. The tip of a germ tube<br />

can penetrate any part of the cuticle. Clavate or<br />

globose appressoria develop at the tips of the<br />

germ tubes. After penetrating the cuticle and<br />

epidermis, the hyphal tips branch and fragment<br />

<strong>to</strong> form multinucleate pro<strong>to</strong>plasts which become<br />

rapidly dispersed throughout the haemocoel<br />

within about 12 24 h. The pro<strong>to</strong>plasts may<br />

increase in number by budding (Butt et al.,<br />

1981). It is believed that the switch <strong>to</strong> the<br />

pro<strong>to</strong>plast form is in response <strong>to</strong> contact with<br />

the nutrient-rich haemolymph. Later, as the<br />

nutrients in the haemolymph are exhausted,<br />

the pro<strong>to</strong>plasts develop cell walls and are<br />

transformed in<strong>to</strong> hyphal bodies. Pro<strong>to</strong>plasts<br />

and hyphal bodies colonize fat bodies, nerve<br />

ganglia and muscle tissue. Infected aphids die<br />

some 72 h after inoculation, and shortly before<br />

death rhizoids develop from enlarged hyphal<br />

bodies and emerge from the ventral side of the<br />

abdomen, making contact with the leaf on which<br />

the aphid has been feeding, then branching by<br />

bifurcation <strong>to</strong> form digitate holdfasts. About<br />

15 30 rhizoids may develop from a single aphid<br />

before it dies. Soon after death, pseudocystidia<br />

and conidiophores emerge. Under natural conditions,<br />

infected aphids die in the late afternoon<br />

and sporulation begins at night. The moist<br />

conditions and dew formation after sunset play<br />

a role in enhancing spore production and<br />

discharge. Hemmati et al. (2001) found that<br />

concentrations of air-borne conidia among<br />

wheat crops were usually highest at night and<br />

in the early morning and relatively low during<br />

the day, peak concentrations being correlated<br />

with high relative humidity.<br />

It is rare for an infected aphid <strong>to</strong> produce<br />

both conidia and resting bodies. Germinating<br />

resting bodies form branched or unbranched<br />

germ tubes bearing retraction septa. They are<br />

terminated by an apical conidium, followed by<br />

one or two lateral conidia. These conidia closely<br />

resemble those which develop on infected<br />

aphids, and they are discharged by septal eversion<br />

(Tyrell & MacLeod, 1975). The germination<br />

of resting bodies is markedly stimulated by longday<br />

conditions of more than 14 h of light per<br />

day (Wallace et al., 1976). In the pea aphid,<br />

E. neoaphidis does not form resting bodies, surviving<br />

instead as hyphal bodies in aphid cadavers.<br />

Artificially infected cadavers can be s<strong>to</strong>red and


ENTOMOPHTHORALES<br />

213<br />

Fig 7.41 Erynianeoaphidis.(a)Hyphalbodyfrom<br />

within an infected aphid. (b) Pointed<br />

pseudocystidium projecting above a layer of<br />

conidiophores at the surface of a dead aphid.<br />

(c) Direct germination of a secondary conidium.<br />

Cy<strong>to</strong>plasm is confined by a retraction septum <strong>to</strong><br />

the tip of the germ tube. (d) Branched<br />

conidiophore with terminal primary conidia.<br />

The wall surrounding the conidium is bitunicate<br />

with a thin outer envelope. (e) Discharged<br />

uninucleate primary conidium. (f) Discharged<br />

secondary conidium.Compare its more ovoid<br />

shape with the shape of the primary conidium.<br />

(g) A discharged secondary conidium has<br />

germinated by repetition <strong>to</strong> form a further<br />

conidium of the same type. Note the bulging<br />

septum on the empty conidium and at the<br />

base of the newly developed conidium.<br />

(h) Primary conidium germinating <strong>to</strong> form<br />

a secondary conidium. Both are bitunicate.<br />

(i) Secondary conidium germinating <strong>to</strong> form<br />

a tertiary conidium with a single large lipid body.<br />

(j) Tertiary conidium germinating by<br />

repetition. (a,b) ¼ 20 mm; (c j) ¼ 12.5 mm.<br />

used as inoculum <strong>to</strong> introduce the parasite<br />

in<strong>to</strong> field populations of aphids pathogenic <strong>to</strong><br />

crops, such as Aphis fabae, the common blackfly<br />

of broad beans. It is also possible <strong>to</strong> grow<br />

E. neoaphidis in agar culture and <strong>to</strong> introduce<br />

inoculum in<strong>to</strong> aphid-infested crops in this form<br />

(Shah et al., 2000).<br />

Erynia conica<br />

Whilst E. neoaphidis shows some versatility in its<br />

asexual reproduction, a more extreme example<br />

is E. conica, which forms four distinct types<br />

of conidium, some of them primary and some<br />

secondary (Descals et al., 1981; Hywel-Jones &<br />

Webster, 1986a). Erynia conica is a parasite of the<br />

blackfly Simulium (Diptera) and some other insect<br />

hosts associated with aquatic habitats. Simulium<br />

spp. have aquatic larval stages generally found in<br />

rapidly flowing streams. The larvae are attached<br />

<strong>to</strong> s<strong>to</strong>nes, twigs and aquatic plants, feeding by<br />

the ingestion of particulate matter collected<br />

by modified branched mouthparts (head rakes).<br />

After pupation, winged adults emerge, mate, and<br />

the females take a blood meal from a mammal.<br />

Gravid females lay egg masses amongst algae and<br />

mosses on water-splashed boulders kept continuously<br />

wet by trickling water. At such sites the<br />

white swollen bodies of dead females infected<br />

by E. conica, attached by rhizoids, may sometimes<br />

be found in large numbers. Conidiophores<br />

project from the carcass, bearing conidia of two<br />

types. From conidiophores which develop in air,<br />

i.e. at the surface of the insect’s body projecting<br />

out of water, boat-shaped bitunicate conidia


214 ZYGOMYCOTA<br />

develop. These are termed primary cornute (type 1)<br />

conidia and are illustrated in Fig. 7.42a. They are<br />

discharged from the conidiophores by septal<br />

eversion. If a type 1 conidium floats on the<br />

water surface, it will germinate <strong>to</strong> produce a<br />

balloon-shaped secondary globose (type 2) conidium<br />

on its upper side, projecting in<strong>to</strong> the air.<br />

The type 2 conidium contains a single large lipid<br />

body and a two-ply septum capable of discharge<br />

by eversion (see Fig. 7.42b).<br />

When primary cornute conidia become<br />

submerged in water, they germinate <strong>to</strong> produce<br />

a secondary conidium of a different type. Its body<br />

has four branches (i.e. it is tetraradiate) and the<br />

position of attachment of this conidium <strong>to</strong> the<br />

short conidiophore is the central point from<br />

where the four arms radiate (see Fig. 7.42c).<br />

This type of spore is termed a secondary stellate<br />

(type 3) conidium. If a dead infected insect is<br />

continuously bathed in water or is submerged,<br />

the conidiophores emerging from it will develop<br />

primary conidia which are also tetraradiate,<br />

but these are attached at the tip of the main<br />

arm from which three upper arms radiate<br />

(see Fig. 7.42d). This type of conidium is a primary<br />

coronate (type 4) conidium. So types 1 and 2 are<br />

aerial conidia, formed and discharged in<strong>to</strong> air,<br />

whilst types 3 and 4 are aquatic conidia, formed<br />

and released under water. Tetraradiate conidia<br />

are a typical adaptation of fungi <strong>to</strong> dispersal in<br />

aquatic environments and are produced also by<br />

aquatic hyphomycetes (see p. 685).<br />

Most of the four types of conidium can<br />

germinate by repetition, by germinating <strong>to</strong><br />

form conidia of one of the other types, or by<br />

the formation of germ tubes. For example, a type<br />

1 conidium can germinate by repetition <strong>to</strong><br />

form a secondary conidium morphologically<br />

identical <strong>to</strong> itself, i.e. another type 1 conidium.<br />

This is described in shorthand as 1 1<br />

Fig 7.42 Erynia conica.The four types of conidia. (a) Primary cornute conidium (type1). Note the bitunicate wall. (b) Primary<br />

cornute conidium germinating <strong>to</strong> produce a secondary globose conidium (type 2). (c) Secondary stellate conidium (type 3) which has<br />

developed from a submerged primary cornute conidium.The point of attachment of the conidium is between the three backwardly<br />

directed arms (arrow). (d) Primary coronate conidium (type 4) with the point of attachment at the end of the main, vertical, arm<br />

(arrow).The single large nucleus is visible below the point of branching.Bar ¼ 20 mm, all images <strong>to</strong> same scale.From Webster (1992),<br />

with kind permission of Springer Science and Business media.


ENTOMOPHTHORALES<br />

215<br />

germination (Webster et al., 1978). Type 1 conidia<br />

may show 1 1, 1 2 and 1 3 germination. Of the<br />

16 (i.e. 4 4) possible interconversions, 12 have<br />

been observed so far (Webster, 1987). The only<br />

type of conidium shown <strong>to</strong> be infective is the<br />

secondary globose, i.e. type 2, conidium (Hywel-<br />

Jones & Webster, 1986b). It is sometimes termed<br />

an invasive conidium. This kind of conidium<br />

only develops from cornute conidia, although<br />

these may be primary or secondary. When a<br />

secondary globose conidium is in contact with an<br />

insect cuticle, a short germ tube develops with<br />

an appressorium at its tip. Penetration of the<br />

cuticle seems <strong>to</strong> be mainly by enzymatic means<br />

and is followed by the formation of multinucleate,<br />

branched hyphal bodies in the haemocoel.<br />

It is presumed that the other kinds of<br />

conidia function as dispersal rather than infection<br />

units, and they can be found in appreciable<br />

numbers trapped in foam near infected flies. It<br />

appears that only adult flies are infected through<br />

the cuticle. Although all types of conidia are<br />

known <strong>to</strong> be present in larval guts, there is no<br />

evidence that larvae are infected from ingested<br />

conidia. Survival over winter, when adult insects<br />

are not available, is by globose, thick-walled<br />

zygospores which are formed within the dead<br />

body of an insect, surrounded by a network of<br />

brown hyphae.<br />

The precise physical conditions associated<br />

with the different types of germination in<br />

E. conica are not known and most attention has<br />

been devoted <strong>to</strong> the germination of the primary<br />

cornute (type 1) and secondary globose (type 2)<br />

conidia (Nadeau et al., 1995, 1996). Germination<br />

of the latter, resulting in appressorium formation<br />

and cuticular penetration on wings of the<br />

susceptible host S. rostratum, occurs over the<br />

temperature range of 15 25°C with an optimum<br />

at 20°C. Germination occurs within 2 h and<br />

penetration within 9 h. The development of<br />

appressoria is related <strong>to</strong> the presence of a coating<br />

of lipid on the host cuticle. In experiments in<br />

which lipids were removed from susceptible<br />

blackfly wings, there was no discernible appressorium<br />

formation or cuticular penetration. On<br />

the non-susceptible host S. decorum, germination<br />

is delayed and appressorium formation and<br />

cuticular penetration do not occur. Instead, a<br />

high level (26%) germination of the 2 1 type<br />

takes place.<br />

The plasticity of asexual reproduction shown<br />

by E. conica is not unique. Similar versatility<br />

is shown by some other members of the<br />

En<strong>to</strong>mophthoraceae which grow on insects<br />

with aquatic larval stages such s<strong>to</strong>neflies<br />

(Plecoptera) and crane flies (Tipulidae) (Descals<br />

& Webster, 1984).<br />

En<strong>to</strong>mophthora muscae<br />

There are about a dozen species of En<strong>to</strong>mophthora,<br />

occurring as widespread insect pathogens<br />

(Samson et al., 1988; Humber, 1997). They are<br />

characterized by unbranched conidiophores and<br />

multinucleate primary conidia which are<br />

projected by a squirt mechanism. Secondary<br />

conidia may form on germination of the primary<br />

conidia, but these are discharged by a septal<br />

eversion mechanism similar <strong>to</strong> that described<br />

above for Erynia and Conidiobolus. Sexual reproduction<br />

is by the formation of zygospores and<br />

azygospores.<br />

The best-known taxon is E. muscae which is, in<br />

fact, a complex of about five species with similar<br />

morphology and spore dimensions (MacLeod<br />

et al., 1976). This fungus is a parasite of houseflies<br />

and other Diptera. Disease is apparent in summer<br />

<strong>to</strong> autumn and is more frequent in wet weather.<br />

In the field, epizootics occur in places where there<br />

are dense populations of potential hosts, for<br />

example dung flies (Sca<strong>to</strong>phaga spp.) on farms, or<br />

hoverflies (Melanos<strong>to</strong>ma spp.) attracted <strong>to</strong> the<br />

honeydew secreted by Claviceps (see Fig. 12.26b)<br />

on the moor grass Molinia. Diseased flies can<br />

occasionally be found attached <strong>to</strong> the glass of a<br />

window pane surrounded by a white halo about<br />

2 cm in diameter made up of discharged conidia<br />

(Plate 3g).<br />

The dead fly shows a distended abdomen with<br />

white bands of conidiophores projecting between<br />

the segments of the exoskele<strong>to</strong>n. The unbranched<br />

multinucleate conidiophores arise from the<br />

coenocytic mycelium which plugs the body of<br />

the dead fly. The conidia are also multinucleate<br />

(Fig. 7.43b). They are projected by a forwardly<br />

directed jet of cy<strong>to</strong>plasm from the elastic conidiophores.<br />

On impact, the bitunicate nature of<br />

the wall of the primary conidium becomes


216 ZYGOMYCOTA<br />

apparent (Fig. 7.43e). Recently discharged conidia<br />

have a dried out drop around them which<br />

represents the cy<strong>to</strong>plasm squirted from the<br />

conidiophore (Figs. 7.43d f). This cy<strong>to</strong>plasmic<br />

coating may act as a protective agent against<br />

desiccation and may possibly help in attaching<br />

the primary conidium <strong>to</strong> the cuticle of an insect.<br />

If the conidium impinges on the body of a fly, it<br />

develops an adhesive pad which attaches it firmly<br />

<strong>to</strong> the cuticle (Fig. 7.43h). Penetration of the<br />

cuticle is probably brought about by a combination<br />

of mechanical and enzymatic means (Brobyn<br />

& Wilding, 1983). A few hours after infection, triradiate<br />

fissures can be seen in the cuticle beneath<br />

attached conidia. When the cuticle in such a<br />

region is examined from the inside, a thin-walled<br />

bladder-like expansion can be seen. From this cell<br />

mycelial branches develop. The hyphae grow<br />

<strong>to</strong>wards the fatty tissues, and as these are<br />

consumed the hyphae break up <strong>to</strong> form wall-less<br />

pro<strong>to</strong>plasts which are carried by the circula<strong>to</strong>ry<br />

system <strong>to</strong> all parts of the body. Eventually the<br />

pro<strong>to</strong>plasts secrete walls and become converted<br />

in<strong>to</strong> hyphal bodies (Fig. 7.43c). Infected flies show<br />

behavioural changes, often crawling <strong>to</strong> the <strong>to</strong>p of<br />

a grass stem and clasping it or adhering <strong>to</strong> walls<br />

or window panes by the proboscis (Maitland,<br />

1994). The sexual behaviour of the host may also<br />

be affected (Moller, 1993). Males attempting <strong>to</strong><br />

mate with diseased females may themselves<br />

Fig 7.43 En<strong>to</strong>mophthora muscae.<br />

(a) House fly adhering <strong>to</strong> a window pane,<br />

surrounded by a halo of discharged<br />

conidia. (b) L.S. house fly showing palisade<br />

of unbranched conidiophores projecting<br />

between segments of the exoskele<strong>to</strong>n.<br />

The conidiophores and conidia are<br />

multinucleate. (c) Hyphal bodies from<br />

recently dead fly extending <strong>to</strong> form<br />

conidiophores. (d) Primary conidium<br />

immediately after discharge surrounded<br />

by cy<strong>to</strong>plasm from the conidiophore.<br />

(e,f) Germination of primary conidia <strong>to</strong><br />

form secondary conidia which are<br />

discharged by bouncing off (septal<br />

eversion). (g) Germination of secondary<br />

conidium by germ tubes. (h) Attachment<br />

of primary conidium <strong>to</strong> integument<br />

of a fly. (i) Two primary conidia<br />

attached <strong>to</strong> integument and penetrating<br />

it by a tri-radiate fissure. (j) View of<br />

penetration from within the integument.<br />

Note the bladder-like expansion within<br />

the tri-radiate fissure. (b g) <strong>to</strong> same<br />

scale, (h j) <strong>to</strong> same scale.


GLOMALES<br />

217<br />

become infected with E. muscae, making it a<br />

sexually transmitted pathogen. Although it was<br />

previously generally accepted that there are no<br />

rhizoids in E. muscae, Balazy (1984) has shown that<br />

rhizoids do develop from hyphal bodies within<br />

the head, growing through the proboscis and<br />

forming a network of branched hyphae with<br />

short irregular holdfasts. A few days after infection<br />

the fly dies and the hyphal bodies within<br />

the abdomen then grow out in<strong>to</strong> coenocytic<br />

hyphae which penetrate between the abdominal<br />

segments and develop in<strong>to</strong> conidiophores.<br />

Discharge of primary conidia begins within<br />

about 5 h, reaching a maximum about 10 12 h<br />

after death. Over 8000 conidia may develop from<br />

a single cadaver (Mullens & Rodriguez, 1985).<br />

The primary conidia remain viable for only<br />

3 5 days. If they fail <strong>to</strong> penetrate a fly, they may<br />

produce secondary conidia within 3 h. The<br />

secondary conidia are released from the tips of<br />

short conidiophores by septal eversion. They may<br />

germinate by a germ tube or by producing the<br />

same type of conidium by repetition.<br />

Within the body of the dead fly, multinucleate<br />

spherical resting bodies (azygospores) are formed.<br />

In the wheat bulb fly Lep<strong>to</strong>hylemia coarctata it has<br />

been observed that a much higher proportion of<br />

infected female flies contain resting bodies as<br />

compared with infected males. This is probably<br />

associated with the longer lifespan of females<br />

than males (Wilding & Lauckner, 1974). Resting<br />

bodies may develop terminally or in an intercalary<br />

position from short hyphae, or by budding<br />

from hyphal bodies. They germinate by developing<br />

a germ conidiophore. Germination is stimulated<br />

by the action of chitin-decomposing<br />

bacteria on the resting spore wall. It is from<br />

such resting bodies that infection probably<br />

begins each year (Goldstein, 1923).<br />

The onion fly Delia antiqua has maggots which<br />

pupate in the soil and overwinter there. Adults<br />

become infected as they emerge through the<br />

soil the following season, presumably from germ<br />

conidia which develop from resting spores<br />

(Carruthers et al., 1985). In some members of<br />

the En<strong>to</strong>mophthorales, e.g. En<strong>to</strong>mophthora sepulchralis,<br />

zygospores develop following conjugation<br />

between hyphal bodies (see Fig. 7.44).<br />

En<strong>to</strong>mophthora muscae, like many other<br />

en<strong>to</strong>mopathogenic En<strong>to</strong>mophthorales, can be<br />

grown in complex media such as those used in<br />

tissue culture (Wolf, 1981; Papierok & Hajek,<br />

1997). Yeast extract and ingredients of animal<br />

origin such as egg yolk, fat and serum or blood<br />

are also used. Growth is markedly stimulated by<br />

glucosamine, a breakdown product of chitin.<br />

Successful cultures have also been established<br />

on a medium containing wheat grain extract,<br />

pep<strong>to</strong>ne, yeast extract and glycerol (Srinivasan<br />

et al., 1964).<br />

Furia<br />

Some of the species formerly placed in the<br />

genus En<strong>to</strong>mophthora have features distinct<br />

from E. muscae and have been re-classified in<strong>to</strong><br />

different genera. An example is Furia americana<br />

(Plate 3h, Fig. 7.45), a fungus found on blowflies<br />

in the autumn, especially around corpses of dead<br />

animals or stinkhorns. In wet weather severe<br />

epidemics may occur, greatly affecting the<br />

blowfly population. Distinctive features are the<br />

conidiophores which branch close <strong>to</strong> the conidiogenous<br />

cells; uninucleate, bitunicate clavate<br />

conidia with a rounded apex and basal papilla;<br />

and discharge by septal eversion. Dead flies are<br />

often attached <strong>to</strong> adjacent plants by filamen<strong>to</strong>us<br />

rhizoid-like hyphae. The conidiophores form<br />

yellowish pustules between the abdominal<br />

segments and the branched tips bear conidia.<br />

The two layers of the conidium wall are<br />

frequently separated from each other by liquid<br />

(Figs. 7.45a c). These conidia are projected for<br />

several centimetres from the host and, on germination,<br />

may form germ tubes, or may produce<br />

secondary conidia which are projected by the<br />

rounding off of a two-ply septum. Within the dry<br />

body of the dead fly numerous smooth hyaline<br />

thick-walled resting spores (azygospores) are<br />

formed by budding from the lateral walls of<br />

parent hyphae (Fig. 7.45f).<br />

7.6 Glomales<br />

The roots of most terrestrial plants grow in<br />

a mutualistic symbiosis with fungi, i.e. an<br />

association in which both partners benefit.


218 ZYGOMYCOTA<br />

mycorrhiza (AM). These fungi are particularly<br />

well-known as mycorrhizal associates of herbaceous<br />

plants, but they may also associate with<br />

trees, especially in the tropics.<br />

Fig 7.4 4 En<strong>to</strong>mophthora sepulchralis.Three stages in<br />

zygospore formation.Two hyphal bodies conjugate and the<br />

zygospore arises as a bud from the fusion cell (afterThaxter,<br />

1888).<br />

Such symbiotic associations are termed mycorrhiza<br />

(Gr. ‘fungus root’). There are several<br />

different kinds of mycorrhiza, including vesicular<br />

and arbuscular mycorrhiza, ec<strong>to</strong>mycorrhiza<br />

(sheathing mycorrhiza, pp. 21 and 526), ericoid<br />

mycorrhiza (p. 442), and orchid mycorrhiza<br />

(p. 596) (Smith & Read, 1997; Peterson et al.,<br />

2004). It is important <strong>to</strong> realize that the nature of<br />

the relationships between the fungi and their<br />

host plants in these distinct types of association<br />

is not the same. In this section we shall look<br />

at the Glomales, a group of zygomyce<strong>to</strong>us<br />

fungi causing the development of vesicular<br />

arbuscular mycorrhiza (VAM) and arbuscular<br />

7.6.1 General features of VAM and AM<br />

A coarse, intercellular, aseptate coenocytic mycelium<br />

within the root tissues may develop large,<br />

balloon-shaped intercalary or terminal thickwalled<br />

vesicles (intraradical vesicles) which are<br />

multinucleate and contain large amounts of<br />

lipid (Figs. 7.46c,d). In some plants, e.g. the<br />

roots of Paris, the mycelium emits branches<br />

which penetrate the cortical root cells, forming<br />

extensive intracellular coils. More commonly,<br />

hyphae penetrating host cells fork repeatedly <strong>to</strong><br />

form richly branched arbuscules (Fig. 7.46c)<br />

which invaginate the plasmalemma. Plant and<br />

fungal plasma membranes are separated by an<br />

apoplastic compartment, the periarbuscular<br />

space. The arbuscule is therefore a type of<br />

haus<strong>to</strong>rium, and there is an interchange of<br />

nutrients and water across the periarbuscular<br />

space. Arbuscules have a relatively short active<br />

life, lasting only a few days. After this time the<br />

fine tips of the arbuscules are digested by the<br />

host cell so that only irregular clumps of fungal<br />

material remain (Fig. 7.46c).<br />

A coarse, angular and often thick-walled<br />

mycelium extends outwards from infected<br />

roots, sometimes for several cm, and penetrates<br />

in<strong>to</strong> the surrounding soil. It may bear large<br />

(4100 mm dia.) globose multinucleate thickwalled<br />

spores which are sometimes termed<br />

chlamydospores. These spores contain thousands<br />

of nuclei as well as energy reserves including<br />

lipid droplets, glycogen, protein and trehalose.<br />

These spores may be borne singly or in clusters<br />

and are often naked, but in some species,<br />

e.g. Glomus mosseae, they are enveloped in a weft<br />

of hyphae <strong>to</strong> form a sporocarp (Figs. 7.46a,b).<br />

Chlamydospores are asexual reproductive structures<br />

and are known <strong>to</strong> survive in dry soil for<br />

many years. For most members of the group only<br />

asexual reproduction is known, but in Gigaspora<br />

decipiens zygospores and azygospores have been<br />

reported in addition <strong>to</strong> chlamydospores.<br />

This species is heterothallic (Tommerup &


GLOMALES<br />

219<br />

Fig 7.45 Furia americana from blowfly.<br />

(a) Branched conidiophore. (b) Single<br />

conidiophore and conidium. Note that the<br />

wall of the conidium is bitunicate. (c) Conidia<br />

after discharge. (d) Conidium germinating by<br />

means of a germ tube. (e) Primary conidia<br />

germinating <strong>to</strong> produce secondary conidia.<br />

(f) Spherical resting bodies from dead fly.<br />

Sivasithamparam, 1990). Sporocarps may form<br />

part of the diet of some mammals and chlamydospores<br />

can be dispersed by soil animals, including<br />

invertebrates and some rodents as well as<br />

larger hoofed mammals. Chlamydospores can<br />

survive in their faeces and are also dispersed in<br />

wind-borne soil dust (Allen, 1991; Allen et al.,<br />

1997; Linderman, 1997).<br />

The spores of Glomales can be extracted from<br />

soil by wet sieving and decanting from soil<br />

slurries using a series of sieves in the<br />

2000 60 mm size range (Gerdemann & Nicolson,<br />

1963). After surface sterilization, single chlamydospores<br />

placed near the roots of susceptible host<br />

plants such as Trifolium and Sorghum germinate,<br />

produce hyphae which make contact with the<br />

root surface and form appressoria before infecting<br />

the root (Hepper, 1984; Menge, 1984). In this<br />

way, dual cultures have been established and can<br />

be maintained by the addition of freshly<br />

extracted spores or infected root pieces <strong>to</strong> pots<br />

containing a suitable host plant. Viable spores<br />

derived from dual cultures maintained on potted<br />

plants are available from the International<br />

Collection of Vesicular Arbuscular Mycorrhizal<br />

<strong>Fungi</strong> (Mor<strong>to</strong>n et al., 1993). Spores have also been<br />

produced under aseptic conditions in association<br />

with hairy root cultures (Mugnier & Mosse, 1987;<br />

Bécard & Fortin, 1988). Limited extension of germ<br />

tubes takes place after germination in vitro, but<br />

sustained growth in the absence of living root<br />

tissues does not occur, so the fungi causing this


220 ZYGOMYCOTA<br />

Fig 7.4 6 Vesicular arbuscular mycorrhiza. (a) Glomus mosseae sporocarp in which chlamydospores are embedded.There are also<br />

naked chlamydospores attached <strong>to</strong> external hyphae. (b) Chlamydospores dissected from a sporocarp, borne on single subtended<br />

hyphae. (c) Onion root cells infected with Glomus mosseae.The cell <strong>to</strong> the left contains a nucleus with two nucleoli and a branched<br />

haus<strong>to</strong>rium or arbuscule. In the cell <strong>to</strong> the right the arbuscule has degenerated. (d) Vesicles from roots of Arum maculatum.


GLOMALES<br />

221<br />

kind of mycorrhiza are obligate mutualistic<br />

symbionts. Glomalean fungi are generally nonspecific<br />

in their host range. The roots of most<br />

groups of vascular land plants are associated with<br />

this type of mycorrhiza, as are the game<strong>to</strong>phytic<br />

stages of Bryophyta and Pteridophyta.<br />

7.6.2 Taxonomy and evolution of Glomales<br />

<strong>Fungi</strong> in this group were originally classified in<br />

the Endogonales but are currently placed in a<br />

separate order, the Glomales, with the<br />

Endogonales now reduced <strong>to</strong> a single genus,<br />

Endogone, with subterranean fleshy sporocarps<br />

which contain zygospores. Each zygospore is<br />

formed after conjugation of two gametangia<br />

(Pegler et al., 1993). One species, E. flammicorona,<br />

forms ec<strong>to</strong>mycorrhizae with some Pinaceae (Fassi<br />

et al., 1969). The fruit bodies of Endogone spp. are<br />

colloquially called ‘pea truffles’ (Plate 3i; Pegler<br />

et al., 1993).<br />

The order Glomales was proposed by Mor<strong>to</strong>n<br />

and Benny (1990) <strong>to</strong> include all soil-borne fungi<br />

which form arbuscules in obligate mutualistic<br />

associations with terrestrial plants. Sexual reproduction<br />

is rare. There are about 150 species and<br />

6 genera in 2 suborders, the Glomineae with<br />

2 families (Glomaceae and Acaulosporaceae),<br />

and the Gigasporineae with a single family<br />

(Gigasporaceae). Members of the Glomineae<br />

(such as Glomus, Acaulospora) form intraradical<br />

vesicles (VAM type), whilst members of the<br />

Gigasporineae have no intraradical vesicles and<br />

are of the AM type. The separation of genera<br />

within the Glomales is based partly on different<br />

patterns of chlamydospore development, and<br />

partly on the structure of the spore wall which<br />

may be complex and multilayered (Hall, 1984;<br />

Mor<strong>to</strong>n & Bentivenga, 1994). Schüssler et al.<br />

(2001) have suggested that the Glomales are not<br />

closely related <strong>to</strong> the Zygomycota and should<br />

be considered as a separate phylum, the<br />

Glomeromycota.<br />

VAM and AM associations are very ancient,<br />

and structures resembling extant arbuscules<br />

have been discovered in the fossilized rhizome<br />

tissues of early vascular plants, including<br />

Devonian psilophytes such as Rhynia (Pirozynski<br />

& Dalpé, 1989; Taylor et al., 1995). Even older are<br />

the fossilized chlamydospores found among<br />

bryophytes of the Ordovician period (some<br />

460 million years old; Redecker et al., 2000a). It<br />

is believed that the origin and evolution of land<br />

plants was dependent on symbiotic associations<br />

of the VAM and AM type (Pirozynski & Malloch,<br />

1975; Malloch, 1987; Simon et al., 1993).<br />

An interesting non-mycorrhizal relative of<br />

the Glomales is the fungus Geosiphon pyriforme,<br />

which is unusual in harbouring a mutualistic<br />

endosymbiont, the cyanobacterium Nos<strong>to</strong>c. When<br />

the hyphal tip of Geosiphon encounters a suitable<br />

symbiont, this is taken up and the hypha<br />

swells <strong>to</strong> form a so-called bladder cell. The<br />

Geosiphon Nos<strong>to</strong>c symbiosis resembles cyanolichens<br />

(see p. 451) in being au<strong>to</strong>trophic both for<br />

carbon and nitrogen (Schüssler & Kluge, 2001).<br />

Geosiphon reproduces by forming chlamydospores<br />

similar <strong>to</strong> those of Glomus. Molecular studies<br />

show that Geosiphon is closely related <strong>to</strong> the<br />

Glomales and may be ancestral <strong>to</strong> the group<br />

(Gehrig et al., 1996; Redecker et al., 2000b).<br />

7.6.3 Physiological and ecological studies<br />

The immense current interest in AM and VAM<br />

mycorrhiza has its origins in the demonstration<br />

of the improved growth of mycorrhiza-infected<br />

host plants compared <strong>to</strong> uninfected controls.<br />

Literature on the physiology of this relationship<br />

has been reviewed by Hause and Fester (2005).<br />

The arbuscule is the main interface for nutrient<br />

exchange between the plant and its fungal<br />

partner, although the latter may also be able <strong>to</strong><br />

take up nutrients through intercellular hyphae.<br />

The periarbuscular space is a highly acidic<br />

compartment (Guttenberger, 2000) due <strong>to</strong> the<br />

outward-directed pumping of pro<strong>to</strong>ns by H þ<br />

ATPases located in the plasma membranes of<br />

both partners. This sets up pro<strong>to</strong>n gradients<br />

which may be used for active uptake of sucrose<br />

hydrolysis products (fruc<strong>to</strong>se and glucose) by the<br />

fungus, and phosphate and other mineral nutrients<br />

by the plant. Pro<strong>to</strong>n-dependent transport<br />

proteins have been localized in both plant and<br />

fungal perihaus<strong>to</strong>rial membranes (see Hause &<br />

Fester, 2005).<br />

The ecology of VAM and AM fungi in crop<br />

plants and natural communities is of particular


222 ZYGOMYCOTA<br />

interest (Allen, 1991; Smith & Read, 1997; Leake<br />

et al., 2004). There are numerous reports of<br />

significant improvements in growth rate, dry<br />

weight and mineral content following infection<br />

especially of plants growing on nutrient-deficient<br />

soils. Emphasis has been placed on phosphate<br />

nutrition. The supply of phosphate (as<br />

2<br />

HPO 4 or H 2 PO 4 , depending on soil pH) is often a<br />

limiting fac<strong>to</strong>r <strong>to</strong> plants growing in natural soils.<br />

It is usually present in low concentrations and<br />

diffuses through soil very slowly. Its influx may<br />

increase 3 4-fold in infected plant roots but<br />

there are also significant increases in other<br />

minerals such as Zn, Cu, and ammonium. The<br />

water relations and resistance of infected plants<br />

<strong>to</strong> infections by pathogens may also be improved.<br />

Increased uptake of minerals is largely due <strong>to</strong> the<br />

exploration of larger volumes of soil by the<br />

extramatrical hyphae which can extend beyond<br />

the depletion zone surrounding plant roots.<br />

The depletion zone is a region in which minerals<br />

are taken up by plant roots at a rate greater than<br />

can be replenished by diffusion through the soil.<br />

For many plants the depletion zone is only<br />

1 2 mm wide whilst the extramatrical hyphae<br />

may extend for several centimetres and can<br />

penetrate in<strong>to</strong> soil cavities <strong>to</strong>o fine <strong>to</strong> be<br />

explored by roots. Moreover, phosphate can be<br />

translocated through fungal hyphae <strong>to</strong>wards<br />

the host root at much faster rates than is possible<br />

by diffusion through soil. The improved growth<br />

of host plants associated with increased supply<br />

of minerals obtained through the hyphae of the<br />

mycorrhizal symbiont is achieved at a cost <strong>to</strong><br />

the plant, i.e. the drain of pho<strong>to</strong>synthate taken<br />

up by the fungus whose biomass, achieved<br />

largely at the expense of the host, may amount<br />

<strong>to</strong> 3 20% of the root weight. In experiments in<br />

which 14 CO 2 was supplied <strong>to</strong> the shoots of young<br />

cucumber plants infected by G. fasciculatum, as<br />

much as 20% of the radioactive carbon fixed by<br />

the plant was used by the fungus (Jakobsen &<br />

Rosendahl, 1990).<br />

In nutrient-deficient soils such as sand dunes,<br />

recently disturbed soil, spoil heaps, areas covered<br />

by volcanic ash, etc., successful colonization by<br />

plants appears <strong>to</strong> be correlated with root infection<br />

by Glomales (Allen, 1991). In closed vegetation<br />

such as mature grassland which contains a<br />

diversity of plants, spore extraction reveals a<br />

wide diversity of AM and VAM species. The roots<br />

of different plant species making up the community<br />

are in close contact and may also be<br />

connected by a hyphal network (Newman,<br />

1988). There is experimental evidence using<br />

iso<strong>to</strong>pically labelled 15 N, 32 P, and 14 C that there<br />

may be an interchange of mineral nutrients and<br />

carbon between unrelated plant species<br />

mediated by VAM mycelia, but Newman (1988)<br />

has cautioned against the conclusion that any<br />

increases in labelled materials necessarily imply<br />

net gains <strong>to</strong> receiver plants at the expense of<br />

donors. There is also experimental evidence<br />

using soil microcosms seeded with a mixture of<br />

grassland grasses and dicotyledons and inoculated<br />

with Glomus constrictum that mycorrhizal<br />

infection may increase species diversity by<br />

selectively enhancing the performance of less<br />

dominant dicotyledons. This results largely<br />

from a reduction in relative abundance of<br />

canopy dominants such as Festuca ovina (Grime<br />

et al., 1987).<br />

7.7 Trichomycetes<br />

The Trichomycetes are a group of fungi which<br />

grow commensally in the guts of terrestrial,<br />

freshwater and marine arthropods such as<br />

insects, millipedes and crustaceans. In most<br />

cases there is little evidence that the host is<br />

harmed by their presence, although it has been<br />

shown that some species may extend parasitically<br />

in<strong>to</strong> the ovarian tissue <strong>to</strong> form chlamydospores<br />

(cysts) in place of eggs. These are deposited<br />

amongst egg masses laid by uninfected females.<br />

McCreadie et al. (2005) have documented an<br />

element of plasticity in the association of a<br />

given trichomycete species, Smittium culisetae,<br />

with its blackfly host which may vary from<br />

commensalistic in well-fed larvae <strong>to</strong> mutualistic<br />

under starvation conditions <strong>to</strong> parasitic if the<br />

ovaries of adult females are infected.<br />

More than 50 genera and over 200 species<br />

have been described but doubtless many more<br />

await discovery (Lichtwardt, 1986, 1996; Misra,<br />

1998; Misra & Lichtwardt, 2000). Members of the


TRICHOMYCETES<br />

223<br />

group have a worldwide distribution and are<br />

especially common in the guts of larvae of<br />

aquatic insects. A few species belonging <strong>to</strong> one<br />

order (Harpellales) have been grown in culture<br />

and appear <strong>to</strong> have no unusual nutritional<br />

requirements. The term trichomycete (Gr. ‘hairy<br />

fungus’) refers <strong>to</strong> the fuzzy appearance of heavily<br />

infested gut linings. Branched or unbranched<br />

thalli are attached by a holdfast <strong>to</strong> the<br />

hindgut cuticle or <strong>to</strong> the peritrophic membrane,<br />

a transparent membranous sleeve which<br />

surrounds digested food material in the mid-gut<br />

of certain insects. Asexual reproduction is by<br />

various types of spore, including trichospores,<br />

chlamydospores, arthrospores or sporangiospores.<br />

Sexual reproduction by the formation<br />

of zygospores is known in the Harpellales.<br />

The occurrence of zygospores, the presence of<br />

chitin in the walls of Smittium culisetae (Sangar &<br />

Dugan, 1973) and molecular studies (O’Donnell<br />

et al., 1998; Gottlieb & Lichtwardt, 2001) all<br />

provide evidence linking Trichomycetes with<br />

the Zygomycota. It is possible that the class<br />

Trichomycetes is polyphyletic, and it is therefore<br />

preferable <strong>to</strong> refer <strong>to</strong> the gut fungi as a biological<br />

group, trichomycetes with a lower-case ‘t’<br />

(Lichtwardt, 1986).<br />

Three orders have been distinguished, namely<br />

the Harpellales, Asellariales and Eccrinales, of<br />

which we shall consider only the first. The<br />

Amoebidiales, previously included, are now<br />

classified with the pro<strong>to</strong>zoa.<br />

7.7.1 Harpellales<br />

Harpella melusinae is one of the most common and<br />

abundant trichomycetes with a worldwide distribution<br />

in temperate regions. It is found in larval<br />

blackflies (Simulium spp.) which live attached <strong>to</strong><br />

s<strong>to</strong>nes, twigs and aquatic vegetation submerged<br />

in rapidly flowing streams. The dissection of<br />

larval guts reveals the peritrophic membrane, <strong>to</strong><br />

the inner wall of which unbranched cylindrical<br />

thalli are attached. Developing thalli receive<br />

nutrients from the material passing through<br />

the gut. The peritrophic membrane is continuously<br />

secreted by endothelial cells lining the<br />

upper part of the mid-gut, i.e. new membrane<br />

material is added at the upper end. Young thalli<br />

are present here and progressively older thalli<br />

further down. Attachment is by a simple holdfast<br />

(Fig. 7.47a). The holdfast consists of a chamber at<br />

the base of the thallus from which numerous<br />

finger-like projections protrude, cemented <strong>to</strong> but<br />

not penetrating the peritrophic membrane<br />

(Reichle & Lichtwardt, 1972). The cylindrical<br />

part of the thallus is divided by septa in<strong>to</strong> 2 12<br />

or more uninucleate segments called generative<br />

cells. The septal ultrastructure consists of a<br />

flared pore associated with a plug, somewhat<br />

resembling the bordered pit of a conifer xylem<br />

tracheid. This feature is characteristic of trichomycetes<br />

(Moss, 1975).<br />

Asexual reproduction<br />

The entire contents of the thallus are converted<br />

<strong>to</strong> reproductive cells. Reproduction begins at the<br />

terminal generative cell and progresses basipetally<br />

(<strong>to</strong>wards the holdfast) by production of<br />

trichospores (Figs. 7.47a c). Trichospores are<br />

really monosporous sporangia. They have been<br />

defined as ‘exogenous, deciduous sporangia<br />

containing a single uninucleate sporangiospore<br />

and normally having one <strong>to</strong> several basally<br />

attached filamen<strong>to</strong>us appendages’ (Lichtwardt,<br />

1986). The trichospores, which are usually coiled<br />

but sometimes straight, develop at the upper end<br />

of a generative cell (Fig. 7.47a). The nucleus of the<br />

generative cell divides mi<strong>to</strong>tically, one daughter<br />

nucleus remaining in the generative cell, the<br />

other entering the developing trichospore. In<br />

H. melusinae there are four basal appendages<br />

which, before trichospore release, are spirally<br />

coiled inside the upper part of the generative cell<br />

(Figs. 7.47b,c; Reichle & Lichtwardt, 1972). At the<br />

distal end of the trichospore within the cy<strong>to</strong>plasm<br />

is an elongated apical spore body<br />

(Fig. 7.47a). This contains holdfast material<br />

which is released after extrusion of the sporangiospore<br />

on its germination within the insect<br />

gut, cementing the holdfast <strong>to</strong> the gut wall (Moss<br />

& Lichtwardt, 1976; Horn, 1989a). Trichospores<br />

are separated from the generative cell by a<br />

septum and are released by breakdown of the<br />

wall beneath it. After release, the appendages<br />

uncoil and extend up <strong>to</strong> 10 times their original<br />

length. The released trichospores are passed out<br />

from the larval gut with faecal material and the


224 ZYGOMYCOTA<br />

Fig 7.47 Harpella melusinae. (a) Thallus<br />

attached by a holdfast <strong>to</strong> the peritrophic<br />

membrane from the mid-gut of a larva of a<br />

blackfly, Simulium sp.The thallus is divided by<br />

septa in<strong>to</strong> three generative cells, each of which<br />

is producing a curved trichospore at its upper<br />

end.The apical spore body at the upper end of<br />

the trichospore contains material extruded <strong>to</strong><br />

cement the holdfast <strong>to</strong> the peritrophic<br />

membrane. (b) Detached trichospore bearing<br />

four filamen<strong>to</strong>us appendages at its base.<br />

(c) Developing trichospore showing the coiled<br />

filamen<strong>to</strong>us appendages wrapped around inside<br />

the wall of the generative cell.<br />

(d) Chlamydospore which has germinated <strong>to</strong><br />

form two generative cells, each of which is<br />

developing a trichospore. (e) Zygospore<br />

development.Two adjacent thalli have<br />

conjugated and from one of the conjugating<br />

cells a zygosporophore has been produced,<br />

terminating in a biconical zygospore. Bars: (a,b)<br />

¼ 10 mm, (c,e) ¼ 5 mm, (d) ¼ 15 mm. (d) after<br />

Moss and Descals (1986); (e) after Lichtwardt<br />

(1967).<br />

appendages cause the trichospores <strong>to</strong> be<br />

entangled with faeces and other particulate<br />

material. Further development of trichospores,<br />

i.e. the extrusion of a sporangiospore from its<br />

sporangium and the formation of a holdfast,<br />

only occurs after ingestion and is stimulated by<br />

conditions in the larval gut.<br />

Smittium culisetae (Harpellales) inhabits the<br />

midgut and hindgut of larval mosqui<strong>to</strong>es and<br />

can be grown in culture. Horn (1989a,b, 1990) has<br />

investigated in vitro the conditions which trigger<br />

sporangiospore extrusion in this and related<br />

species of Smittium. The trigger for extrusion in<br />

S. culisetae is a two-stage process. Phase 1, which<br />

simulates mid-gut conditions, involves exposure<br />

<strong>to</strong> 20 mM KCl at pH 10 followed by phase 2, in<br />

which the pH is reduced <strong>to</strong> 7, simulating hindgut<br />

conditions. Following this sequence of treatments<br />

sporangiospores (¼ trichospores sensu<br />

stric<strong>to</strong>) are rapidly extruded, a process in which<br />

they increase in size by the uptake of water<br />

and by vacuolation, generating turgor pressure<br />

which aids extrusion. Spore germination quickly<br />

follows with the secretion of holdfast material<br />

from the apical spore body through canals in the<br />

distal wall of the sporangiospore (Horn, 1989a).<br />

A second asexual, free-living stage in the<br />

life cycle of H. melusinae is the chlamydospore<br />

(sometimes termed ovarian cyst or cys<strong>to</strong>spore),<br />

masses of which are deposited by adult female


TRICHOMYCETES<br />

225<br />

Simulium in the place of eggs (Moss & Descals,<br />

1986; Lichtwardt, 1996). A similar stage has<br />

been reported from Simulium infected with<br />

Genistellospora homothallica (Labeyrie et al., 1996).<br />

Ovarian tissue is invaded by the fungus, growing<br />

in a parasitic mode. Cysts of H. melusinae dissected<br />

from ovaries are surrounded by a membranous<br />

sheath. They are ellipsoidal and, on germination,<br />

form two germ tubes, one at each pole, ending as<br />

a spherical knob, the generative cell initial. A<br />

single generative cell develops from each initial<br />

and forms a terminal trichospore (Fig. 7.47d). The<br />

chlamydospores are deposited among egg masses<br />

and infection of young larvae results from<br />

ingestion of trichospores produced by them. The<br />

ovarian chlamydospores therefore represent<br />

a ‘missing link’ in the life cycle of Harpellales.<br />

Adult blackflies do not contain trichomycete<br />

thalli because at the final ecdysis (moult)<br />

before pupation, the cuticular lining of the<br />

larval gut is shed.<br />

Sexual reproduction<br />

This occurs by the production of zygospores and<br />

has so far been reported only in Harpellales. In<br />

H. melusinae zygospores are rarely detected,<br />

possibly because they are associated with the<br />

last stage of development of the Simulium larval<br />

host before pupation and are shed at ecdysis.<br />

Zygospore formation is preceded by conjugation<br />

between swollen cells on adjacent thalli (see<br />

Fig. 7.47e; Lichtwardt, 1967). From one of the<br />

conjugating cells a zygosporophore grows out,<br />

and from this a biconical zygospore develops.<br />

The biconical shape, which is characteristic of<br />

the Harpellales, is possibly adapted <strong>to</strong> passage<br />

through the insect gut. In some members of the<br />

group zygospores bear polar filamen<strong>to</strong>us appendages,<br />

but these are absent in Harpella (Moss &<br />

Lichtwardt, 1977). The cy<strong>to</strong>logical details of<br />

zygospore formation have not been fully worked<br />

out but in H. melusinae the zygospore, zygosporophore<br />

and the two conjugant cells each contain<br />

a single nucleus. Moss and Lichtwardt (1977) have<br />

speculated that the four nuclei might have been<br />

derived from meiotic division of a diploid zygote<br />

nucleus within the fused conjugants. On this<br />

hypothesis the conjugant cells would be interpreted<br />

as gametangia, a situation markedly<br />

different from that found in other zygomycetes.<br />

Relationships<br />

On the basis of similarities in serological reactions,<br />

septal pore structure and in sporangial<br />

morphology, it has been suggested that the<br />

Harpellales are related <strong>to</strong> the Kickxellales, an<br />

order of mostly dung- and soil-inhabiting saprotrophic<br />

zygomycetes (Moss & Young, 1978).<br />

However, the evidence for a phylogenetic relationship<br />

between these two groups is conflicting.<br />

K.L. O’Donnell et al. (1998) and Gottlieb and<br />

Lichtwardt (2001) have attempted <strong>to</strong> correlate<br />

morphological criteria with molecular data (18S<br />

rDNA) but found only poor support, whereas<br />

Tanabe et al. (2004), comparing a range of DNA<br />

sequences, found a strong link between<br />

Harpellales and Kickxellales (see Fig. 7.1).


8<br />

Ascomycota (ascomycetes)<br />

8.1 <strong>Introduction</strong><br />

The phylum Ascomycota (colloquially called<br />

ascomycetes) is by far the largest group of<br />

fungi, estimated <strong>to</strong> include more than 32 000<br />

described species in 3400 genera (Kirk et al.,<br />

2001). It is assumed that the majority of<br />

ascomycetes has yet <strong>to</strong> be discovered, and the<br />

<strong>to</strong>tal number of species may well be higher by<br />

a fac<strong>to</strong>r of 10 20 or even more (see Hawksworth,<br />

2001). The name is derived from the Greek words<br />

askos (a leather bottle, bag or bladder) and mykes<br />

(a fungus), so ascomycetes are sac fungi. The<br />

characteristic feature of the group is that the<br />

sexually produced spores, the ascospores<br />

(see p. 25), are contained within a sac, the<br />

ascus. In most ascomycetes the ascus contains<br />

eight ascospores and is turgid, ejecting its spores<br />

by a squirt mechanism.<br />

There is a very wide range of lifestyles.<br />

Some ascomycetes are saprotrophs, others<br />

necrotrophic or biotrophic parasites of plants<br />

and animals, including humans. Examples of<br />

biotrophic parasites are the Erysiphales, the<br />

cause of many powdery mildew diseases of<br />

plants (Chapter 13), the Taphrinales (p. 251)<br />

causing a range of plant diseases associated<br />

with growth abnormalities, and the<br />

Laboulbeniales, relatively harmless ec<strong>to</strong>parasites<br />

of beetles and some other insects (Blackwell,<br />

1994; Weir & Blackwell, 2001). Many ascomycetes<br />

grow as endophytes in symp<strong>to</strong>mless associations<br />

with plants. Some are mutualistic symbionts, for<br />

example the lichens (Chapter 16) which make up<br />

about 40% of the described species of ascomycetes.<br />

Lichens are dual organisms consisting of<br />

a fungus (usually an ascomycete) and a pho<strong>to</strong>synthetic<br />

alga or cyanobacterium living in close<br />

association. This type of association has evolved<br />

independently in several unrelated groups of<br />

ascomycetes and indeed it has been claimed that<br />

several major fungal lineages are derived from<br />

lichen-symbiotic ances<strong>to</strong>rs (Lutzoni et al., 2001),<br />

although this hypothesis is under dispute (Liu &<br />

Hall, 2004; see Fig. 8.17). Symbiotic mycorrhizal<br />

relationships also exist between true truffles<br />

(e.g. Tuber spp.) or false truffles (e.g. Elaphomyces<br />

spp.) and trees such as oak and beech (see pp. 423<br />

and 313). The range of habitats is wide, as would<br />

be expected of such a large and diverse group of<br />

fungi. Ascomycetes grow in soil, are common on<br />

the above-ground parts of plants, and are also<br />

found in freshwater and in the sea.<br />

Most ascomycetes are recognized by their<br />

fruit bodies or ascocarps, i.e. the structures<br />

which surround the asci. These will be described<br />

more fully later (see Fig. 8.16).<br />

8.2 Vegetative structures<br />

Ascomycetes may grow either as yeasts, i.e.<br />

unicells multiplying by budding or fission,<br />

or as mycelia consisting of septate hyphae<br />

(Fig. 8.1a). Some fungi may switch from the<br />

yeast <strong>to</strong> the filamen<strong>to</strong>us state or vice versa,<br />

i.e. they are dimorphic. A good example of a<br />

dimorphic fungus is Candida (see Fig. 8.1b).


VEGETATIVE STRUCTURES<br />

227<br />

Fig 8.1 (a) Hypha of Hormonema dematioides.The positions of<br />

the first septa are indicated by arrows. (b) Candida parapsilosis.<br />

Pseudohypha budding off cells which continue <strong>to</strong> bud in a<br />

yeast-like manner.<br />

Candida albicans is the cause of diseases such as<br />

thrush in mammals, including man.<br />

The mycelial septa of ascomycetes are usually<br />

incomplete, developing as transverse centripetal<br />

flange-like ingrowths from the cylindrical wall of<br />

a hypha, which fail <strong>to</strong> meet at the centre so that<br />

in most ascomycete septa there is one central<br />

pore permitting cy<strong>to</strong>plasmic continuity and<br />

streaming between adjacent segments of mycelium<br />

(Buller, 1933; Gull, 1978). This means that<br />

organelles such as mi<strong>to</strong>chondria and nuclei<br />

are free <strong>to</strong> travel from cell <strong>to</strong> cell; the large<br />

nuclei are constricted as they pass through the<br />

pore (Fig. 8.2). Individual cells may be uni- or<br />

multinucleate and the cy<strong>to</strong>plasmic continuity<br />

between the cells means that the mycelium<br />

of an ascomycete is effectively coenocytic.<br />

Proteinaceous organelles termed Woronin<br />

bodies (Buller, 1933) may be closely grouped<br />

near the central pore (Fig. 8.3). Woronin bodies<br />

are globose structures or ‘hexagonal’ (polyhedral)<br />

crystals made up essentially of one protein<br />

(Tenney et al., 2000), and surrounded by a unit<br />

membrane. They measure 150 500 nm in width<br />

and are sufficiently large <strong>to</strong> block the septal<br />

pore. They rapidly do so near regions where<br />

a hypha is physically damaged. Usually one<br />

Woronin body blocks one pore. The blockage in<br />

the septal pore is consolidated by deposition of<br />

further material (for references see Markham &<br />

Collinge, 1987; Markham, 1994; Momany et al.,<br />

2002). Woronin bodies are formed near the<br />

hyphal apex and are transported <strong>to</strong> more distal<br />

regions of the hypha as septa develop. Woronin<br />

bodies have been recorded from ascomycetes and<br />

their related conidial fungi, but there are no<br />

reliable reports from other fungal phyla.<br />

The mycelium of many ascomycetes is<br />

homokaryotic (Gr. homoios ¼ like, resembling;<br />

karyon ¼ a nut, meaning nucleus), i.e. all nuclei<br />

in a given mycelium are genetically identical.<br />

Heterokaryotic mycelia also occur and generally<br />

arise through anas<strong>to</strong>mosis, i.e. the cy<strong>to</strong>plasmic<br />

fusion of vegetative hyphae. Following anas<strong>to</strong>mosis<br />

between homokaryons of differing genotypes,<br />

nuclei, other organelles and plasmids may<br />

be transferred between one mycelium and<br />

another so that a given mycelium or even a<br />

single cell may contain nuclei of different kinds.<br />

However, the ability <strong>to</strong> form heterokaryons is<br />

under genetic control and a degree of genetic<br />

similarity between homokaryons is necessary<br />

for it <strong>to</strong> occur. Failure <strong>to</strong> establish a heterokaryon<br />

is a phenomenon known as heterokaryon<br />

incompatibility or vegetative incompatibility<br />

(Caten & Jinks, 1966; see pp. 320 and 594).


228 ASCOMYCOTA (ASCOMYCETES)<br />

Fig 8.2 Nucleus of Botrytiscinerea passing through a septal<br />

pore. (a) View of the entire hyphal diameter. (b) Close up.Note<br />

the constricted appearance of the nucleus.<br />

Fig 8.3 Transmission electron micrograph of a transverse<br />

septum in the hypha of Emericella nidulans showing<br />

five Woronin bodies near the central septal pore.<br />

Scale bar ¼ 0.25 mm. Reprinted from Momany et al. (2002),<br />

Mycologia,withpermission.ßThe Mycological Society of<br />

America.<br />

Some ascospores, e.g. those of Neurospora tetrasperma,<br />

are heterokaryotic, and multinucleate<br />

conidia can also be heterokaryotic. Following<br />

mycelial anas<strong>to</strong>mosis between homokaryons,<br />

nuclear division succeeded by migration may<br />

result in the rapid spread of an introduced<br />

nucleus in<strong>to</strong> a mycelium, thus transforming a<br />

homokaryon in<strong>to</strong> a heterokaryon. An ascomycete<br />

mycelium may thus consist of a mosaic of cells,<br />

some of which are homokaryotic and others<br />

heterokaryotic. Because the different types of<br />

nuclei do not always divide at the same rate,<br />

the ratio of nuclear types in a heterokaryotic<br />

mycelium may change with time and respond <strong>to</strong><br />

changes in external conditions such as nutrient<br />

availability. This gives the mycelium a degree of<br />

genetic flexibility which sometimes manifests<br />

itself as the formation of sec<strong>to</strong>rs in a mycelium<br />

in agar culture (Fig. 8.4). Another important<br />

source of genetic variability which may arise<br />

within a heterokaryon is the parasexual cycle, a<br />

process in which genetic recombination is<br />

brought about in the absence of meiosis. This is<br />

discussed more fully below.<br />

8.3 Life cycles of ascomycetes<br />

8.3.1 Sexual life cycles<br />

Sexual life cycles in the strict sense, i.e. involving<br />

nuclear fusion and meiosis, occur only in those<br />

ascomycetes which possess asci, because it is<br />

within the young ascus that these events occur.


LIFE CYCLES OF ASCOMYCETES<br />

229<br />

Fig 8.4 Sec<strong>to</strong>ring of a mycelial colony of Pseudeurotium sp.<br />

The slow-growing wild-type mycelium appears dark due <strong>to</strong> the<br />

formation of melanized cleis<strong>to</strong>thecia.Two non-fruiting sec<strong>to</strong>rs<br />

have formed which show faster vegetative growth but no<br />

sporulation on the rich agar medium.<br />

Ascospores of most ascomycetes contain one or<br />

more haploid nuclei, and therefore most (but by<br />

no means all) ascomycetes have a haploid<br />

vegetative mycelium. The mycelium is often<br />

capable of asexual reproduction, e.g. by fragmentation,<br />

budding or by the formation of conidia,<br />

chlamydospores, sclerotia, etc. The structure and<br />

formation of conidia is described below. Some<br />

yeasts, e.g. Saccharomyces cerevisiae, show an alternation<br />

of diploid and haploid yeast-like states<br />

and here the diploid state is the commonly<br />

encountered form (p. 265), in contrast <strong>to</strong><br />

Schizosaccharomyces in which the vegetative cells<br />

are haploid (p. 253).<br />

The mating behaviour of ascomycetes may<br />

be homothallic or heterothallic. In homothallic<br />

ascomycetes the mycelium derived from a single<br />

ascospore is capable of reproducing sexually,<br />

i.e. by developing asci. Examples are Emericella<br />

nidulans, Pyronema domesticum and Sordaria fimicola.<br />

However, the homothallic condition does<br />

not preclude outcrossing as is shown by the<br />

formation of hybrid asci containing black<br />

(wild type) and white (mutant) ascospores in<br />

crosses between different strains of Sordaria<br />

fimicola (see Fig. 12.2). In heterothallic ascomycetes<br />

the ascus usually contains four ascospores<br />

of one mating type and four of the other.<br />

The two mating types differ at a single allele<br />

and the mating types may be designated A and a,<br />

a and a, or (þ) and ( ). Sexual reproduction<br />

occurs following plasmogamy between cells of<br />

the two mating types. Plasmogamy is of three<br />

main types:<br />

1. Gametangio-gametangiogamy. Fusion occurs<br />

between differentiated gametangia. An example<br />

is Pyronema domesticum where fusion<br />

is between the trichogyne, a filamen<strong>to</strong>us<br />

extension of the large, swollen ‘female’<br />

gametangium (the ascogonium) and a<br />

less swollen ‘male’ gametangium, the antheridium,<br />

which donates nuclei <strong>to</strong> the trichogyne<br />

and thereby <strong>to</strong> the ascogonium (see<br />

p. 416).<br />

2. Game<strong>to</strong>-gametangiogamy. Fusion takes place<br />

between a small unicellular male gamete<br />

(spermatium) and a differentiated female<br />

gametangium (ascogonium). The spermatium<br />

is rarely capable of independent germination<br />

and growth and may only germinate <strong>to</strong><br />

produce a short conjugation tube which<br />

fuses with the wall of the ascogonium. An<br />

example is Neurospora crassa in which the<br />

spermatium fuses with a trichogyne (see<br />

Fig. 12.7).<br />

3. Soma<strong>to</strong>gamy. Fusion takes place between<br />

undifferentiated hyphae, i.e. there are no<br />

recognizable sexual organs. This type of<br />

sexual behaviour is shown by Coprobia granulata,<br />

whose orange ascocarps are common on<br />

cattle dung.<br />

8.3.2 Asexual life cycles<br />

Most fungi which were formerly classified in the<br />

artificial group Deuteromycotina or <strong>Fungi</strong><br />

Imperfecti are conidial forms (anamorphs) of<br />

Ascomycota, although a few have affinities with<br />

Basidiomycota. Evidence for a relationship <strong>to</strong><br />

Ascomycota comes from morphological similarity<br />

and from DNA sequence comparisons.<br />

Morphological similarities include the structure<br />

of the mycelium, the layering of the hyphal


230 ASCOMYCOTA (ASCOMYCETES)<br />

wall as seen by electron microscopy, the finestructure<br />

of nuclear division, and also close<br />

resemblances of conidial structure and development.<br />

Some genera contain species which reproduce<br />

by asexual means only, whilst closely<br />

similar forms have sexual as well as asexual<br />

reproduction. Examples include Aspergillus and<br />

Penicillium, which are anamorphs of several<br />

genera of Ascomycota (Trichocomaceae; see<br />

pp. 308 313) and Fusarium which is the<br />

anamorph of Gibberella and Nectria (members of<br />

the Hypocreales; see p. 343). It is presumed that<br />

fungi which reproduce only by conidia have lost<br />

the capacity <strong>to</strong> form ascocarps in the course of<br />

evolution.<br />

8.3.3 Parasexual reproduction<br />

This is a process in which genetic recombination<br />

can occur through nuclear fusion and crossingover<br />

of chromosomes during mi<strong>to</strong>sis. Meiosis<br />

does not occur, and instead haploidization takes<br />

place by the successive loss of chromosomes<br />

during mi<strong>to</strong>tic divisions. It is believed that the<br />

necessary cy<strong>to</strong>logical steps take place in a regular<br />

sequence which Pontecorvo (1956) has termed<br />

the parasexual cycle. The essential steps include<br />

(i) nuclear fusion between genetically distinct<br />

haploid nuclei in a heterokaryon <strong>to</strong> form<br />

diploid nuclei; (ii) multiplication of the diploid<br />

nuclei along with the original haploid nuclei;<br />

(iii) the development of a diploid homokaryon;<br />

(iv) genetic recombination by crossing-over<br />

during mi<strong>to</strong>sis in some of the diploid nuclei;<br />

and (v) haploidization of some of the diploid<br />

nuclei by progressive loss of chromosomes<br />

(aneuploidy) during mi<strong>to</strong>sis.<br />

This process was discovered in Emericella<br />

(Aspergillus) nidulans, which can reproduce sexually<br />

by forming asci and asexually by forming<br />

conidia (see Fig. 11.17). By changing the nutrient<br />

content of the medium on which the fungus is<br />

grown, the development of asci and therefore of<br />

normal sexual reproduction can be prevented.<br />

Genetic mapping based on gene recombination<br />

following conventional sexual reproduction<br />

has been compared with mapping based on<br />

parasexual recombination and has yielded identical<br />

results.<br />

Parasexual recombination is known <strong>to</strong> occur<br />

not only in Ascomycota but also in Oomycota<br />

and Basidiomycota. It makes possible genetic<br />

recombination in organisms not known <strong>to</strong> reproduce<br />

by sexual means and helps us <strong>to</strong> understand<br />

why purely asexual fungi such as many<br />

species of Aspergillus and Penicillium have achieved<br />

success and have continued <strong>to</strong> flourish in the<br />

course of evolution. However, because parasexual<br />

reproduction is comparatively rare in nature,<br />

it is probably only a partial substitute for sexual<br />

reproduction, so that purely asexual species are<br />

more prone <strong>to</strong> accumulating deleterious mutations<br />

(Geiser et al., 1996).<br />

8.4 Conidia of ascomycetes<br />

The asexual spores or conidia of ascomycetes<br />

are remarkably diverse in form, structure and<br />

modes of dispersal, but their development or<br />

conidiogenesis occurs in a limited number of<br />

ways (see below). The cell from which a conidium<br />

develops is the conidiogenous cell and usually<br />

one or more such cells are borne on a stalk, the<br />

conidiophore. Conidiophores which are narrow<br />

and not differentiated from the vegetative mycelium<br />

are said <strong>to</strong> be micronema<strong>to</strong>us (Gr. nema ¼ a<br />

thread) whilst those that are clearly differentiated<br />

are macronema<strong>to</strong>us. Conidiophores<br />

frequently arise singly as in Eurotium repens<br />

(Fig. 11.16), Emericella nidulans (Fig. 11.17) and in<br />

many species of Penicillium (Fig. 11.18).<br />

However, in certain fungi the conidiophores<br />

may aggregate <strong>to</strong> form a conidioma. Descriptive<br />

terms have been given <strong>to</strong> different types<br />

of conidioma. Conidiophores aggregated in<strong>to</strong><br />

parallel bundles (fascicles) are termed coremia<br />

(Gr. korema ¼ a brush) or synnemata (Gr. prefix<br />

syn ¼ <strong>to</strong>gether). Examples are Penicillium<br />

claviforme (Fig. 11.19) and Cephalotrichum<br />

(Dora<strong>to</strong>myces) stemonitis (Fig. 12.39). Seifert (1985)<br />

has distinguished several types of synnema, some<br />

of which are simple, some compound, some<br />

made up of parallel conidiophores, and others


CONIDIUM PRODUCTION IN ASCOMYCETES<br />

231<br />

where the hyphae making up the synnema are<br />

intricately interwoven (see Kirk et al., 2001).<br />

In many ascomycetes the conidiophores develop<br />

on or in a stroma (Gr. stroma ¼ bed, cushion),<br />

an aggregation of pseudoparenchyma<strong>to</strong>us cells.<br />

A good example of a conidial stroma is seen in<br />

the wood-rotting candle-snuff fungus, Xylaria<br />

hypoxylon (see Fig. 12.11a). Here, powdery white<br />

conidia develop at the tips of the branches of<br />

the conidial stroma and, later, asci develop in<br />

flask-shaped perithecia at the base of the old<br />

stroma. The term sporodochium (Gr. spora ¼ a<br />

seed; doche ¼ a receptacle) is used for the<br />

cushion-like conidiomata bearing a layer of<br />

short conidiophores. An example is the conidial<br />

(Tubercularia) state of Nectria cinnabarina (see<br />

Fig. 12.20c).<br />

Another type of conidioma is the acervulus<br />

(Lat. acervulus ¼ a little heap), a saucer-shaped<br />

fructification which may develop inside the<br />

tissues of a host plant or may be superficial.<br />

Subepidermal acervuli develop from a pseudoparenchyma<strong>to</strong>us<br />

stroma, and as the acervulus<br />

matures the overlying epidermis of the host<br />

becomes ruptured <strong>to</strong> expose conidia formed<br />

from conidiogenous cells lining the base of the<br />

saucer. The conidia are held <strong>to</strong>gether in slime<br />

and are chiefly dispersed by rain splash. A good<br />

example of an acervular fungus is Colle<strong>to</strong>trichum<br />

(see Fig. 12.51). The teleomorphs of Colle<strong>to</strong>trichum,<br />

where known, are species of Glomerella, many of<br />

which are serious plant pathogens. In many<br />

ascomycetes and their allies, the conidia are<br />

borne inside flask-shaped conidiomata termed<br />

pycnidia (Gr. diminutive of pyknos ¼ dense,<br />

packed, concentrated). Traditionally, fungi with<br />

pycnidial and acervular states have been grouped<br />

<strong>to</strong>gether in the artificial taxon coelomycetes<br />

(Sut<strong>to</strong>n, 1980), in contrast <strong>to</strong> hyphomycetes in<br />

which the conidiogenous cells are exposed on<br />

single conidiophores or in synnemata, coremia<br />

or sporodochia (see above). Pycnidia may be<br />

superficial or embedded in host tissue. The opening<br />

of the pycnidium is generally by means of<br />

a circular ostiole. Conidia formed from conidiogenous<br />

cells lining the inner wall of the<br />

pycnidium are held <strong>to</strong>gether in slimy masses<br />

which ooze out through the ostiole, sometimes<br />

as spore tendrils. They are generally dispersed<br />

by splash or in water films. In some cases the<br />

pycnidia, instead of producing conidia with an<br />

asexual function, produce spermatia which are<br />

involved in fertilization. Examples of fungi with<br />

pycnidial anamorphs are Lep<strong>to</strong>sphaeria acuta with<br />

a Phoma anamorph (Fig. 17.3), and Phaeosphaeria<br />

nodorum (anamorph Stagonospora nodorum; see<br />

Fig. 17.4).<br />

8.5 Conidium production<br />

in ascomycetes<br />

There are several steps in the production and<br />

release of conidia, namely (1) conidiogenesis,<br />

i.e. conidial initiation; (2) maturation; (3) delimitation;<br />

(4) secession, i.e. separation from<br />

the conidiogenous cell; (5) proliferation of the<br />

conidiogenous cell or conidiophore <strong>to</strong> form<br />

further conidia. Many of the current ideas on<br />

conidiogenesis stem from a seminal paper by<br />

Hughes (1953) based on light microscopy<br />

studies of conidial development in a range of<br />

hyphomycetes. Hughes classified the development<br />

of conidia in a limited number of ways.<br />

His ideas were extended by other workers, and<br />

advances were also made possible by the use<br />

of electron microscopy and time-lapse cinepho<strong>to</strong>micrography<br />

(Cole & Samson, 1979).<br />

An excellent review of these aspects of<br />

conidiogenesis has been written by Cole (1986).<br />

The descriptions which follow are based on the<br />

account by de Hoog et al. (2000a). Conidiogenesis<br />

occurs in two ways which appear <strong>to</strong> be distinct<br />

at first glance: blastic and thallic (see Fig. 8.5).<br />

In reality, when surveying conidium formation<br />

and release in a range of fungi, there is a<br />

continuum of development of which these two<br />

concepts represent extremes (Minter et al., 1982,<br />

1983a,b; Minter, 1984).<br />

8.5.1 Blastic conidiogenesis<br />

The conidium develops by the blowing-out of<br />

the wall of a cell, usually from the tip of a hypha,<br />

sometimes laterally as in Aureobasidium (conidial


232 ASCOMYCOTA (ASCOMYCETES)<br />

Fig 8.5 Two basic modes of development during<br />

conidiogenesis of a hyphal apex (a): blastic (b) and thallic (c).<br />

From de Hoog et al. (2000a), with kind permission of<br />

Centraalbureau voor Schimmelcultures.<br />

state of Discosphaerina; see Fig. 17.25). In certain<br />

yeasts including Saccharomyces, the blastic development<br />

of new daughter cells is known as<br />

budding (see Fig. 10.3). Two kinds of blastic<br />

development have been distinguished:<br />

1. Holoblastic. All the wall layers of the<br />

conidiogenous cell contribute <strong>to</strong> the wall of<br />

the newly formed conidium (see Fig. 8.6b).<br />

Aureobasidium pullulans (Fig. 17.25) and Tricladium<br />

splendens (conidial Hymenoscyphus; see Fig. 25.12)<br />

are examples. In some genera with dark (i.e.<br />

melanized), relatively thick-walled conidiophores<br />

such as Stemphylium and Alternaria (anamorphs<br />

of Pleospora), the conidia develop holoblastically,<br />

but a narrow channel persists in the wall of<br />

the conidiogenous cell through which cy<strong>to</strong>plasm<br />

had passed as the spore expanded. This type of<br />

development has been described as porogenous<br />

(Luttrell, 1963) or tretic (Ellis, 1971a) and the<br />

conidia are sometimes termed porospores or<br />

poroconidia (see Figs. 17.10 17.13; Carroll &<br />

Carroll, 1971; Ellis, 1971b).<br />

2. Enteroblastic. The wall of the conidiogenous<br />

cell is rigid and breaks open. The initial of<br />

the conidium is pushed through the opening<br />

and is surrounded by a newly formed wall<br />

(Fig. 8.6c). Two types of enteroblastic development<br />

have been distinguished, phialidic and<br />

annellidic (see Fig. 8.7). In phialidic development<br />

a basipetal succession of conidia (phialospores,<br />

phialoconidia) develops from a specialized conidiogenous<br />

cell, the phialide (Gr. diminutive of<br />

phialis ¼ flask), usually shaped like a bottle<br />

with a narrow neck. Phialides are formed singly<br />

Fig 8.6 Two alternative types of conidiogenesis starting from<br />

an undifferentiated hyphal apex (a). In holoblastic<br />

conidiogenesis (b), the entire wallbecomes inflated <strong>to</strong> form the<br />

conidium initial. In enteroblastic conidiogenesis (c), the<br />

conidium initial develops through a hole in the rigid outer wall.<br />

From de Hoog et al. (2000a), with kind permission of<br />

Centraalbureau voor Schimmelcultures.<br />

or in clusters at the tip of a conidiophore or,<br />

more rarely, laterally. There may be one or<br />

several nuclei in a phialide. As shown in Fig. 8.8<br />

for Thielaviopsis basicola, the initial of the firstformed<br />

phialoconidium is surrounded by the<br />

apical wall of the phialide and is, in reality,<br />

holoblastic. The phialide wall breaks transversely<br />

near its tip and the first conidium, surrounded<br />

by a newly formed wall and capped by the wall<br />

from the broken tip of the phialide, is pushed<br />

out (Hawes & Beckett, 1977; Ingold, 1981). The<br />

new wall material which encases the phialoconidium<br />

is secreted in the form of a cylinder from<br />

the surface of the cy<strong>to</strong>plasm deep within the<br />

phialide, a process known as ring wall building<br />

(Minter et al., 1983a). Before the conidium is<br />

extruded, a septum develops within the phialide<br />

below its neck, at the base of the conidium.<br />

The upper part of the wall of the now open-ended<br />

phialide persists as a small collar, the collarette.<br />

The nucleus or nuclei within the phialide<br />

continue <strong>to</strong> divide mi<strong>to</strong>tically. A second conidium<br />

develops below the first, and is surrounded<br />

by newly secreted wall material. This conidium is<br />

also cut off by a septum and pushed out. Part of<br />

the newly secreted wall material may persist<br />

around the inside of the neck of the phialide as<br />

periclinal thickening. The process is repeated<br />

so that many phialoconidia may develop from<br />

a single phialide. In phialides which have<br />

developed several conidia, the periclinal


CONIDIUM PRODUCTION IN ASCOMYCETES<br />

233<br />

Fig 8.7 Enteroblastic phialidic (left) and enteroblastic (right) annellidic conidiogenesis. From de Hoog et al.(2000a),withkind<br />

permission of Centraalbureau voor Schimmelcultures.<br />

Phialidic conidiogenesis<br />

Annellidic conidiogenesis<br />

a. Apex of conidiophore expands <strong>to</strong> form a phialide a. Apex of conidiophore differentiates <strong>to</strong> form a conidiogenous<br />

with a blown-out holoblastic conidium initial (o).<br />

cell (annellide) and the initial of a holoblastic first conidium (1).<br />

b. The first-formed conidium (1), surrounded by a new wall b. The first conidium is cut off by a septum.<br />

secreted inside the phialide, is pushed out and breaks<br />

the outer wall of the phialide whose tip persists as a cap.<br />

c. The first conidium is cut off by a septum. c. A second conidium (2) develops beneath the first.<br />

d. A second conidium develops below the first, also d. The septum cutting off the second conidium is formed beyond<br />

surrounded by new wall secreted inside the phialide. the point at which the original annellide wall was ruptured<br />

and persists as an annellation.<br />

e. A third conidium develops in basipetal succession<br />

adding <strong>to</strong> the length of the conidial chain.The lower<br />

part of the broken original wall of the phialide has<br />

persisted as a collarette, the extent of which is shown<br />

by vertical dashed lines. Successive layers of wall<br />

material may accrete in the neck of the phialide <strong>to</strong><br />

form a periclinal thickening.<br />

e. The development of further conidia results in the addition<br />

of more annellations so that an annellated zone, marked by<br />

vertical dashed lines, increases in length.<br />

thickening may be seen even with the light<br />

microscope, but in others it is less obvious.<br />

During maturation of the phialoconidium,<br />

the spore may increase in size, its wall may<br />

become thickened and ornamented by spines<br />

and may become pigmented by melanin and<br />

other materials. In some genera of ascomycetes<br />

and their conidial derivatives, the phialospores<br />

are dry and appear in chains. Dry-spored conidial<br />

chains are often persistent and are typical of<br />

Aspergillus and Penicillium (see Figs. 11.16 11.18).<br />

The hydrophobic nature of the spore wall is due<br />

<strong>to</strong> incorporation of hydrophobin rodlets. The<br />

adherence of the conidia in chains depends on<br />

the strength of the septum between adjacent<br />

spores. Secession of the conidia in<strong>to</strong> separate<br />

spores occurs by breakage of the septum. Where<br />

the spore wall is wet, the succession of conidia<br />

may briefly persist in the form of a short chain<br />

(false chain) or may collapse in<strong>to</strong> slimy balls at<br />

the tips of the phialides (no-chain phialides).<br />

An example of the latter is Trichoderma (conidial


234 ASCOMYCOTA (ASCOMYCETES)<br />

Fig 8.8 Conidiogenesis inThielaviopsis basicola.<br />

(a) Phialides and phialoconidia.To the left is a<br />

branched conidiophore with a short phialide<br />

which has not yet formed conidia and a longer<br />

phialide in which the tip has broken<br />

transversely and the first phialoconidium is<br />

being extruded.This spore is capped by the<br />

remnants of the phialide tip. Developing<br />

phialoconidia can be seen in the necks of the<br />

phialides. (b) Three end spores from a spore<br />

chain.The capped terminal spore is more<br />

bulbous than the cylindrical spores which<br />

succeedit. (c) A branched conidiophore bearing<br />

two phialides with chains of hyaline thin-walled<br />

phialoconidia and a dark, thick-walled<br />

transversely septate chlamydospore.These<br />

two distinct conidial states are synanamorphs.<br />

Scale bar: (a) ¼ 10 mm, (b) ¼ 20 mm.<br />

Hypocrea; see Fig. 12.16). Phialoconidia are, in<br />

general, unicellular but multicellular conidia<br />

are found in certain genera such as in the transversely<br />

septate conidium of Sporoschisma (conidial<br />

Melanochaeta; for references see Sivichai et al.,<br />

2000).<br />

Annellidic conidiogenesis (Fig. 8.7) in many<br />

ways resembles phialidic, and indeed the term<br />

annellidic phialide is sometimes used for this<br />

type of conidiogenous cell. These are also termed<br />

annellides (Lat. annulus ¼ little ring) or annellophores,<br />

and the spores which develop from them<br />

are annelloconidia. As in phialidic development,<br />

the first-formed annelloconidium is holoblastic.<br />

The difference between the two modes of<br />

development is that new wall material which is<br />

secreted within the annellide protrudes beyond<br />

its neck and the septum which cuts off the newly<br />

formed conidium also forms beyond the neck.<br />

As each new conidium develops in basipetal<br />

fashion, a small ring of wall material (annellation)<br />

is left at the neck of the annellide, which<br />

thus grows in length as successive conidia<br />

develop. This accumulation of short collars of<br />

wall material is the annellated zone (see Fig. 8.7).<br />

With normal light microscopy annellation may<br />

be difficult <strong>to</strong> see, but detection is improved by<br />

interference contrast or phase contrast optics.<br />

Examples of fungi reproducing by annelloconidia<br />

are Scopulariopsis brevicaulis (see Fig. 12.38;<br />

Cole & Kendrick, 1969a) and Cephalotrichum<br />

(Dora<strong>to</strong>myces) stemonitis (Fig. 12.39). Both genera<br />

contain species which are conidial forms of<br />

Microascus.


CONIDIUM PRODUCTION IN ASCOMYCETES<br />

235<br />

Secession of conidia, irrespective of their<br />

mode of development, is in most cases by<br />

dissolution of the septum or septa which<br />

separate them from the conidiogenous cell or<br />

from adjacent spores. This process is termed<br />

schizolytic secession (Gr. schizo ¼ <strong>to</strong> split, divide;<br />

lyticos ¼ able <strong>to</strong> loosen). In some other cases<br />

secession is brought about by the collapse of a<br />

special separating cell beneath the terminal<br />

conidium. This is termed rhexolytic secession<br />

(Gr. rhexis ¼ a rupture, breaking).<br />

8.5.2 Thallic conidiogenesis<br />

Thallic conidiogenesis (Gr. thallos ¼ a branch)<br />

occurs by conversion of a pre-existing hyphal<br />

element in which terminal or intercalary cells of<br />

a hypha become cut off by septa (see Fig. 8.9). Two<br />

kinds of thallic development have been distinguished:<br />

holothallic (Gr. holos ¼ whole, entire)<br />

and thallic-arthric (Gr. arthron ¼ a joint).<br />

In holothallic development a hyphal element,<br />

e.g. a terminal segment of a hypha, is converted<br />

as a whole in<strong>to</strong> a single conidium (see Fig. 8.9).<br />

Secession of such conidia may be schizolytic<br />

or rhexolytic. Microsporum spp. (anamorphic<br />

Arthroderma), which are skin pathogens (derma<strong>to</strong>phytes)<br />

of mammals, provide examples of this<br />

holothallic development (see Fig. 11.6). During<br />

thallic-arthric conidiogenesis, septa develop in<br />

a hypha and divide it up in<strong>to</strong> segments which<br />

separate in<strong>to</strong> individual cells by dissolution of<br />

the septa (see Fig. 8.9). Geotrichum candidum<br />

(anamorphic Galac<strong>to</strong>myces), a common soil<br />

fungus and frequent contaminant of milk and<br />

milk products, develops conidia in this way<br />

(see Fig. 10.10; Cole, 1975).<br />

The proliferation of the conidiogenous cell<br />

or the conidiophore may occur in various ways,<br />

for example by the formation of a new growing<br />

point in the region of the conidiophore beneath<br />

the point at which the first conidium was<br />

formed. The new apex extends beyond the<br />

point of origin of the first conidium and develops<br />

a new conidiogenous cell. These methods of<br />

conidiophore regeneration are discussed more<br />

fully in relation <strong>to</strong> some of the different genera.<br />

Fig 8.9 Holothallic and thallic arthric conidiogenesis.From de Hoog et al. (2000a), with kind permission of Centraalbureau voor<br />

Schimmelcultures.<br />

Holothallic conidiogenesis with rhexolytic secession Thallic arthric conidiogenesis with schizolytic secession<br />

a. The terminal portion of a hypha is cut off by a septum. a. A terminal segment of a hypha.<br />

b. A second septum laid down near the first cuts off a b. Septa develop, dividing the segment in<strong>to</strong> several cells.<br />

subterminal segment, the separating cell.<br />

c. The terminal cell enlarges <strong>to</strong> form the conidium. c. The septa divide, each separating in<strong>to</strong> two layers.<br />

d. Collapse of the separating cell causes conidium secession. d. The daughter cells separate.


236 ASCOMYCOTA (ASCOMYCETES)<br />

8.6 Development of asci<br />

The morphogenesis of asci and ascospores has<br />

been reviewed by Read and Beckett (1996). In<br />

yeasts and related fungi, the ascus arises directly<br />

from a single cell, but in most other ascomycetes<br />

it develops from a specialized hypha, the ascogenous<br />

hypha, which in turn develops from an<br />

ascogonium (Fig. 8.10a). The ascogenous hypha of<br />

many ascomycetes is multinucleate, and its tip<br />

is recurved <strong>to</strong> form a crozier (shepherd’s crook).<br />

Within the ascogenous hypha, nuclear division<br />

occurs simultaneously. Two septa at the tip of<br />

the crozier cut off a terminal uninucleate cell<br />

and a penultimate binucleate cell (Fig. 8.10c)<br />

destined <strong>to</strong> become an ascus. The ante-penultimate<br />

cell beneath the penultimate cell is termed<br />

the stalk cell. The terminal cell of the crozier<br />

curves round and fuses with the stalk cell, and<br />

this region of the ascogenous hypha may grow<br />

on <strong>to</strong> form a new crozier in which the same<br />

sequence of events is repeated. Repeated proliferation<br />

of the tip of the crozier can result in<br />

a tight cluster of asci in many ascomycetes or<br />

a succession of well-separated asci as in Daldinia<br />

concentrica (see Fig. 12.10c). Specialized septal<br />

plugs, more elaborate than normal Woronin<br />

bodies, block the pores in the septa at the base<br />

of the ascus (Kimbrough, 1994). The septal pore<br />

plugs probably aid in retaining the high turgor<br />

pressure which develops in asci shortly before<br />

ascospore discharge.<br />

In the ascus initial the two nuclei fuse and<br />

the diploid fusion nucleus undergoes meiosis <strong>to</strong><br />

form four haploid daughter nuclei (Figs. 8.10d,e).<br />

Fig 8.10 Diagrammatic representation of cy<strong>to</strong>logical features during ascus development. (a) Ascogenous hypha with a crozier at its<br />

tip developing from an ascogonium. (b) Conjugate nuclear division of the two nuclei in the crozier. (c) Two septa have cut off a<br />

binucleate penultimate cell.The two nuclei fuse <strong>to</strong> form a diploid nucleus.The uninucleate terminal segment of the ascogenous hypha<br />

has recurved and fused with the ascogenous hypha <strong>to</strong> form the stalk cell. (d) The penultimate cell enlarges <strong>to</strong> become an ascus initial<br />

within which the fusion nucleus begins <strong>to</strong> divide meiotically. A new crozier is developing from the stalk cell. (e) Second division of<br />

meiosis has occurred in the developing ascus.The behaviour of the new crozier repeats that of the first. (f) Mi<strong>to</strong>tic division of the<br />

four haploid nuclei resulting from meiosis in the first ascus. (g) Ascospores formed.


DEVELOPMENT OF ASCI<br />

237<br />

These nuclei then undergo a mi<strong>to</strong>tic division<br />

so that eight haploid nuclei result (Fig. 8.10f).<br />

The eight nuclei may divide further mi<strong>to</strong>tically<br />

so that each ascospore is binucleate, or,<br />

if still more mi<strong>to</strong>ses follow, the ascospore<br />

becomes multinucleate. For example, a single<br />

mi<strong>to</strong>sis occurs in the immature ascospores of<br />

Neurospora crassa, and the spores remain binucleate<br />

for 2 3 days after they have been delimited.<br />

Later, a series of four or more synchronous<br />

mi<strong>to</strong>ses occur after the spores have become<br />

pigmented so that they contain 32 or more<br />

nuclei when they are mature (Raju, 1992a).<br />

Where the ascospores are multicellular, there<br />

are repeated nuclear divisions accompanied by<br />

the formation of septa which divide up the spore.<br />

In some ascomycetes more than eight ascospores<br />

are formed, usually in numbers which are a<br />

multiple of eight, e.g. in the coprophilous genera<br />

Podospora and Thelebolus. In others the eight<br />

multicellular ascospores break up in<strong>to</strong> partspores,<br />

e.g. in Hypocrea (Fig. 12.15c) and<br />

Cordyceps spp. (Fig. 12.33b). In Taphrina ascospores<br />

may bud mi<strong>to</strong>tically within the ascus so that<br />

the mature ascus contains numerous yeast cells<br />

(Fig. 9.2c). Asci with fewer than eight spores<br />

are also known, e.g. in Neurospora tetrasperma<br />

where the four ascospores are binucleate, in<br />

Phyllactinia guttata where there are two ascospores<br />

(Fig. 13.14e), or in Monosporascus cannonballus<br />

which has a single ascospore.<br />

8.6.1 Cleavage of ascospores<br />

In many ascomycetes, studies of the fine structure<br />

of asci during cleavage of the ascospores<br />

have shown that a system of double membranes<br />

continuous with the endoplasmic reticulum<br />

extends from the envelope of the diploid fusion<br />

nucleus (Fig. 8.11). The double membrane develops<br />

<strong>to</strong> form a cylindrical envelope lining the<br />

young ascus. This peripheral membrane cylinder<br />

or lining layer is termed the ascus vesicle or<br />

ascospore-delimiting membrane. The ascospores<br />

are cut out from the cy<strong>to</strong>plasm within the<br />

ascus by infolding and fusion of the inner<br />

edges of the double membrane around a portion<br />

of cy<strong>to</strong>plasm and a nucleus (Fig. 8.11d). In some<br />

ascomycetes, e.g. Taphrina, a peripheral<br />

membrane cylinder has not been observed and<br />

the nuclei within the ascus become enveloped by<br />

ascospore-delimiting membranes formed by<br />

direct invagination of discrete parts of the<br />

ascus plasma membrane.<br />

Between the two layers of the ascosporedelimiting<br />

membrane enclosing the ascospores,<br />

the primary spore wall is secreted. The inner<br />

membrane forms the plasma membrane of the<br />

ascospore and the outer membrane becomes<br />

the spore-investing membrane. Secondary wall<br />

material is secreted within the primary wall.<br />

There may be several such layers. In Sordaria<br />

humana a <strong>to</strong>tal of four spore wall layers have<br />

been distinguished, a primary wall layer and<br />

three secondary layers (Read & Beckett, 1996).<br />

The secondary wall layers are often quite thick,<br />

and in dark-walled ascospores the pigment is<br />

usually laid down within the secondary wall<br />

layers. The spore wall may be smooth or<br />

extended <strong>to</strong> form a variety of ornamentations<br />

such as spines, ridges or reticulations. The ascus<br />

epiplasm, i.e. the residual cy<strong>to</strong>plasm remaining<br />

outside the spores after these have become<br />

cleaved out, may continue <strong>to</strong> play a part in the<br />

formation of the ascospore wall. For example,<br />

in Ascobolus immersus, the outer leaf of the sporedelimiting<br />

membrane may extend irregularly<br />

outwards in<strong>to</strong> the surrounding epiplasm <strong>to</strong> form<br />

a perisporic sac within which secondary wall<br />

material is deposited, derived from the epiplasm,<br />

and passing as globular bodies through the<br />

membrane of the perisporic sac. This secondary<br />

wall material ornaments the ascospore wall but<br />

is not involved in the formation of the purple<br />

pigment characteristic of Ascobolus ascospores<br />

(Wu & Kimbrough, 1992).<br />

In many ascomycetes the outermost layer of<br />

the ascospore wall, the perispore, is mucilaginous,<br />

as seen for example in Ascobolus immersus<br />

(Fig. 14.5), Sordaria fimicola (Fig. 12.1c) and<br />

Pleospora herbarum (Fig. 17.10a). The properties<br />

of this outer wall layer may aid in the lubrication<br />

of the spore and also enable it <strong>to</strong> be compressed<br />

as it emerges from the ascus. Further, it may aid<br />

in the attachment of ascospores <strong>to</strong> substrata.<br />

It may also cause ascospores <strong>to</strong> stick <strong>to</strong>gether<br />

<strong>to</strong> form multisporous projectiles, an adaptation<br />

which results in an increased distance of


238 ASCOMYCOTA (ASCOMYCETES)<br />

Fig 8.11 Ascus development in Ascobolus (after Oso,1969). (a) Young ascus showing the formation of membrane-bounded vesicles<br />

(V) from the nucleus (N).The ascus wall (AW) is lined by the plasmalemma (PM). (b) Appearance of the ascospore membrane (AM)<br />

at the tip of the ascus and the arrangement of vesicles along the periphery of the ascus. (c) Ascospore membrane now in the form of<br />

a peripheral tube open at the lower end.The diploid nucleus has divided. (d) Invagination of the ascospore membrane between the<br />

haploid nuclei. (e) Young ascospores (S) delimitedby the ascospore membrane from the epiplasm (E). (f) Separation of the two layers<br />

of the ascospore membrane due <strong>to</strong> the formation of the primary spore wall (PSW) between them.<br />

ascospore discharge as compared with singlespored<br />

projectiles (Ingold & Hadland, 1959). This<br />

adaptation is especially common in coprophilous<br />

fungi, i.e. those which grow and fruit on<br />

herbivore dung, such as Ascobolus (Fig. 14.5) and<br />

Sordaria. A special adaptation occurs in another<br />

coprophilous fungus, Podospora, in which the<br />

basal part of the spore proper develops as a<br />

primary appendage, whilst other parts of the<br />

perispore extend as mucilaginous secondary


DEVELOPMENT OF ASCI<br />

239<br />

appendages (see Figs. 12.3, 12.4; Beckett et al.,<br />

1968). The appendages of adjacent spores intertwine<br />

so that the spores are discharged strung<br />

<strong>to</strong>gether in the manner of a slingshot (Ingold,<br />

1971). In some aquatic ascomycetes the ascospores<br />

have extensions of the spore wall which aid<br />

in attachment. Pleospora scirpicola, which forms<br />

ascocarps on the submerged parts of culms<br />

of Schoenoplectus lacustris, an inhabitant of the<br />

shoreline of freshwater lakes, canals and slowmoving<br />

rivers, has long, mucilaginous, tapering<br />

extensions from each end of the ascospore<br />

(Fig. 17.1d). Appendaged ascospores are especially<br />

common in marine ascomycetes. The appendages<br />

develop in a variety of ways and unfurl in sea<br />

water, slowing down their rate of sedimentation<br />

and increasing the likelihood of their attachment<br />

<strong>to</strong> underwater substrata such as wood<br />

(Hyde & Jones, 1989; Hyde et al., 1989; Jones,<br />

1994).<br />

Germ pores or germ slits, through which<br />

germ tubes emerge on spore germination, are<br />

found in many ascomycetes, especially those<br />

with thick dark-pigmented walls. Germ pores,<br />

representing thin areas in the spore wall, occur<br />

at each end of the spore in Neurospora and<br />

germination may occur at either or at both<br />

ends. In Sordaria humana there is a single germ<br />

pore at the lower end of the ascospore plugged<br />

by a pore plug (Read & Beckett, 1996). The<br />

ascospores of Xylariaceae, e.g. Xylaria, Hypoxylon<br />

and Daldinia, have black walls with a hyaline<br />

germ slit running along the length of the spore<br />

(Figs. 12.10 and 12.14).<br />

Because the division which follows the fournucleate<br />

stage is mi<strong>to</strong>tic and because the division<br />

plane is usually parallel <strong>to</strong> the length of the<br />

ascus, adjacent pairs of spores starting from the<br />

tip of an ascus are normally sister spores and<br />

are thus genetically identical. Rare exceptions <strong>to</strong><br />

this situation are occasionally found where the<br />

division planes are oblique, or for other reasons<br />

(see Raju, 1992a).<br />

8.6.2 The ascus wall<br />

The wall of the ascus consists of several<br />

distinguishable layers. The outer layer is laid<br />

down first and inside it is a succession of laterformed<br />

layers so that the mature wall may<br />

consist of four or more layers (Bellemère, 1994;<br />

Read & Beckett, 1996). The wall material<br />

includes chitin, polysaccharides and proteins,<br />

but there is no evidence of lipid. The ascus<br />

wall is elastic. All or parts of it may stretch<br />

considerably during ascospore liberation, and<br />

contraction of the elastic wall provides the force<br />

for ascospore discharge. During discharge all<br />

the layers of the ascus wall may remain<br />

attached <strong>to</strong> each other, thus appearing as a<br />

single layer. Such asci are termed unitunicate<br />

(Lat. tunica ¼ a garment). Despite the term<br />

unitunicate which refers <strong>to</strong> the behaviour<br />

(i.e. function) of the ascus wall during ascus<br />

dehiscence, the wall of unitunicate asci is often<br />

composed of two superposed tunicae, a thin,<br />

single-layered or double-layered exoascus and<br />

a thicker endoascus. The endoascus may be<br />

fibrillar, or at first granular and then with<br />

parallel or reticulate fibrils (Parguey-Leduc<br />

& Janex-Favre, 1984). During ascospore discharge<br />

the two layers of the ascus wall<br />

remain attached, i.e. they do not glide over<br />

each other.<br />

A variant of the unitunicate type of ascus<br />

dehiscence is found in the lichenized ascomycetes<br />

Lecanora and Physcia (the Lecanora or rostrate<br />

type of dehiscence). In Physcia stellaris the ascus<br />

has a prominent amyloid dome. The ripening<br />

ascospores push against this dome and on<br />

ascospore discharge it is extruded <strong>to</strong> form<br />

a rostrum (Lat. rostrum ¼ beak) which extends<br />

upwards <strong>to</strong> the surface, whilst its base remains<br />

attached <strong>to</strong> the upper part of the wall of the<br />

ascus (see Figs. 8.12e,f; Honegger, 1978).<br />

In other ascomycetes, the ascus wall appears<br />

distinctly two-layered (bitunicate) when viewed<br />

with the light microscope (Luttrell, 1951;<br />

Reynolds, 1971, 1989). The layers of many (but<br />

not all) bitunicate asci separate at ascospore<br />

discharge in<strong>to</strong> two functionally distinct layers<br />

(see Fig. 8.12), and such asci are termed fissitunicate<br />

(Dughi, 1956). Fissitunicate asci are particularly<br />

common in the Loculoascomycetes.<br />

Development of the wall of a bitunicate ascus<br />

takes place in two stages prior <strong>to</strong> ascospore<br />

formation. The first stage involves the growth<br />

of the ascus initial and the expansion of the


240 ASCOMYCOTA (ASCOMYCETES)<br />

Fig 8.12 Types of ascus dehiscence. (a) Pro<strong>to</strong>tunicate ascus; the wall dissolves <strong>to</strong> release the ascospores passively. (b,c) Operculate<br />

asci before and after discharge; the ascus opens by means of a lid or operculum. (d) Discharged inoperculate ascus which has opened<br />

through a pore. (e,f) Rostrate ascus as seen in Physcia. In (e) a thickened part of the upper wall of the ascus is being extruded and<br />

is visible in a discharged ascus (f) as an extension of the inner part of the ascus wall, the rostrum. (g) Discharged bilabiate ascus<br />

showing the longitudinal slit by which the ascus opens. (h,i) Bitunicate ascus before and after the first stage of spore release.<br />

Rupture of the ec<strong>to</strong>tunica has allowed the endotunica <strong>to</strong> expand. (e,f) after Honegger (1978).<br />

ascus mother cell. During this stage the outer<br />

layers of the wall making up the ec<strong>to</strong>tunica<br />

(¼ ec<strong>to</strong>ascus) are deposited. In the second stage,<br />

secondary wall layers making up the endotunica<br />

(¼ endoascus) are laid down within the primary<br />

wall. The development of bitunicate asci has<br />

been studied by Reynolds (1971) and by Parguey-<br />

Leduc and Janex-Favre (1982). At the beginning<br />

of development, asci are surrounded by a single<br />

homogeneous layer which is sometimes granular,<br />

bearing externally a loose network (a fuzzy<br />

coat) of interascal material. The ascus wall<br />

becomes divided in<strong>to</strong> a densely granular external<br />

layer (the ec<strong>to</strong>ascus) and a clearer, but equally<br />

granular, inner layer (the endoascus). A clear<br />

space then separates these two layers. The<br />

granular material of the endoascus rearranges<br />

itself in<strong>to</strong> lines of fibrils at first following a wavy<br />

pattern as seen in transverse sections of developing<br />

asci. Later the fibrils become strongly<br />

folded in<strong>to</strong> pointed zigzag shapes, a development<br />

which progresses from the inside <strong>to</strong>wards the<br />

outside of the endoascus. The folds of the zigzags<br />

are closely pressed against the pointed teeth<br />

which mark out the plasmalemma. The density<br />

of the fibrils increases considerably throughout<br />

the thickness of the endoascus. Finally the two<br />

layers of the ascal wall, separated from each<br />

other by a clear space, appear as a double-layered<br />

ec<strong>to</strong>ascus and a single-layered endoascus within<br />

which the fibrils are strongly pleated in<strong>to</strong><br />

accordion-like folds. The pointed crests of the<br />

pleats lie parallel <strong>to</strong> each other and perpendicular<br />

<strong>to</strong> the plasmalemma of the ascus. The<br />

folding of the layers of the endoascus and<br />

plasmalemma permit the rapid expansion of<br />

the ascus prior <strong>to</strong> spore discharge, i.e. by<br />

providing material which can unfold rapidly.<br />

Towards the tip of the non-discharged ascus<br />

the crests of the pleated folds of the endoascus


DEVELOPMENT OF ASCI<br />

241<br />

may converge and appear as a kind of apical<br />

basket which has been termed by Chadefaud<br />

(1942) the nasse apicale (Fr. nasse ¼ keep net,<br />

eel trap).<br />

In some asci the ascus wall does not function<br />

in ascospore discharge, but dissolves or<br />

disintegrates at maturity, the spores being<br />

released passively. Such asci are termed pro<strong>to</strong>tunicate<br />

(see Fig. 8.12a). This type of ascus is<br />

characteristic of certain groups of ascomycetes<br />

such as the Eurotiales and Onygenales but they<br />

are also found in unrelated groups (Currah, 1994).<br />

Examples are Eurotium (Fig. 11.16) and Gymnoascus<br />

(Fig. 11.9).<br />

8.6.3 The apical apparatus of asci<br />

The apical dome of the ascus may be modified in<br />

various ways. In certain types of discomycete<br />

with an open saucer-like fruit body or apothecium,<br />

the ascus is capped by a wall which has<br />

an annulus of thinner wall material forming a<br />

lid or operculum (Lat. operculum ¼ a cover, lid)<br />

(van Brummelen, 1981). When the ascus explodes<br />

<strong>to</strong> discharge its ascospores, the operculum may<br />

be lifted off completely or may hinge <strong>to</strong> one side<br />

(see Figs. 14.5, 14.6). Such asci are operculate<br />

(Figs. 8.12b,c). However, the majority of ascomycetes<br />

have no ascus lid; they are inoperculate<br />

and when ascospore discharge occurs, the tip<br />

of the ascus opens by a pore (Fig. 8.12d).<br />

The presence or absence of an operculum is<br />

a character used in the classification of<br />

discomycetes. Operculate asci are characteristic<br />

of Pezizales including genera such as Aleuria,<br />

Ascobolus and Pyronema (see Chapter 14).<br />

Inoperculate discomycetes include Helotiales<br />

such as Sclerotinia (Figs. 15.1, 15.2) and many<br />

other orders. In a few cases, e.g the lichenized<br />

ascomycete Pertusaria and the coprophilous<br />

fungus Ascozonus, the ascus may burst by one<br />

or two longitudinal slits at the apex (see<br />

Fig. 8.12g). Such asci are described as bilabiate<br />

(i.e. two-lipped).<br />

Other kinds of specialized structures found in<br />

ascus tips are generally referred <strong>to</strong> as the apical<br />

apparatus. Their functions relate <strong>to</strong> the mechanism<br />

of discharge (see below). In many perithecial<br />

fungi the tip of the ascus contains an apical<br />

ring or annulus. This is a specially thickened<br />

inward extension of the apical wall of the ascus,<br />

arranged in the form of a cylindrical flange<br />

(Fig. 8.13). In some fungi (e.g. Xylaria) the annulus<br />

is amyloid, i.e. it stains blue with Melzer’s iodine,<br />

an aqueous solution of I 2 in KI (Beckett &<br />

Crawford, 1973). In other ascomycetes, reddishbrown<br />

(dextrinoid) staining may be observed,<br />

whereas in yet others (e.g. Sordaria) the annulus<br />

does not stain with iodine (Read & Beckett, 1996).<br />

When ascospores are discharged the annulus<br />

is everted, i.e. turned inside out like a sleeve<br />

(see Fig. 8.13b). The cylindrical opening of the<br />

annulus is considerably less than the diameter of<br />

the ascospores which pass through it as shown<br />

in Fig. 8.13a for Xylaria longipes and Fig. 12.1 for<br />

Sordaria fimicola, so that the annulus must be<br />

sufficiently elastic <strong>to</strong> expand and contract as an<br />

ascospore passes through it. The function of the<br />

annulus is <strong>to</strong> act as a sphincter, minimizing the<br />

decrease in hydrostatic pressure inside the ascus<br />

as spore discharge proceeds. It may also separate<br />

the spores from each other as they pass through,<br />

and by gripping the tapering rear portion of an<br />

ascospore impart some force which helps <strong>to</strong><br />

expel it (Ingold, 1954a). Filiform (i.e. needleshaped)<br />

ascospores are discharged singly and<br />

not in groups. This is well shown in the ergot<br />

fungus Claviceps purpurea and its allies such as<br />

Cordyceps (Figs. 12.27, 12.33). Their ascus<br />

apices are capped by a swollen plug of wall<br />

material pierced by a narrow pore. Ascospores<br />

are squeezed out through the pore, sometimes<br />

with an interval of several seconds between<br />

successive discharge events (Ingold, 1971).<br />

Other elaborations of the upper part of the<br />

ascus have been reported (Beckett, 1981;<br />

Bellemère, 1994).<br />

The actual form of the mature ascus is very<br />

variable. In pro<strong>to</strong>tunicate forms, i.e. with nonexplosive<br />

ascospore release, the ascus is often<br />

a globose sac, but in the majority of ascomycetes<br />

the ascus is cylindrical, and the spores are<br />

expelled from the ascus explosively.<br />

8.6.4 Hamathecium<br />

In many cases the asci are surrounded by<br />

packing tissue in the form of paraphyses


242 ASCOMYCOTA (ASCOMYCETES)<br />

Fig 8.13 Xylaria longipes. Fine structure of the ascus apex<br />

(after Beckett & Crawford,1973). (a) L.S. undischarged ascus<br />

showing the apical ring. (b) L.S. discharged ascus showing the<br />

eversion of the apical ring.<br />

(Gr. para- ¼ near, beside, parallel; physis ¼<br />

growth), or pseudoparaphyses. The general<br />

term for such sterile inter-ascal tissue is the<br />

hamathecium (Gr. hama ¼ all <strong>to</strong>gether, at the<br />

same time) (Eriksson, 1981). Paraphyses are filaments<br />

which are attached <strong>to</strong> the ascocarp near<br />

the bases of the asci and are free at their upper<br />

ends as in Pyronema (Figs. 14.2a,c) and Ascobolus<br />

(Fig. 14.6). Pseudoparaphyses are hyphae which<br />

usually arise above the level of the asci and grow<br />

downwards between them. They may become<br />

attached at their lower ends as in Pleospora<br />

(Fig. 17.9). Because the paraphyses and pseudoparaphyses<br />

pack tightly around the asci, the<br />

latter cannot expand laterally but are forced <strong>to</strong><br />

elongate. A hamathecium is lacking in certain<br />

groups of ascomycetes, e.g. the Eurotiales<br />

and Clavicipitales, and also in Mycosphaerella<br />

(Fig. 17.19). The sum of all contents of the<br />

ascoma (i.e. the hamathecium plus asci) but<br />

excluding the ascoma wall is called the centrum.<br />

8.6.5 The mechanism of<br />

ascospore discharge<br />

Explosive release of ascospores follows increased<br />

turgor pressure, caused by water uptake by the<br />

ascus. In the young ascus, after the spores have<br />

been cut out, the epiplasm remains lining the<br />

ascus wall, and this surrounds a large central<br />

vacuole containing ascus sap, within which<br />

the ascospores are suspended. The epiplasm is<br />

rich in the polysaccharide glycogen which can<br />

be visualized cy<strong>to</strong>chemically by its reddishbrown<br />

staining with the I 2 /KI stain. As the ascus<br />

matures, the red stain diminishes in intensity<br />

due <strong>to</strong> the conversion of the polysaccharide <strong>to</strong><br />

osmolytes of lower molecular weight. This brings<br />

about an increased osmotic concentration of<br />

the ascus sap, followed by increased water<br />

uptake. The resulting increase in turgor pressure<br />

causes the ascus <strong>to</strong> stretch and, eventually,<br />

<strong>to</strong> burst open, squirting out the ascospores.<br />

The osmotic pressure of the sap in mature asci<br />

extending from apothecia of Ascobolus immersus<br />

has been determined <strong>to</strong> be up <strong>to</strong> 3 bar (0.3 MPa),<br />

with glycerol being the main organic osmolyte<br />

(Fischer et al., 2004). In Gibberella zeae, the turgor<br />

pressure required for ascus discharge (1.54 MPa)<br />

seems <strong>to</strong> be caused mainly by a K þ and Cl influx<br />

across the plasma membrane, with the most<br />

abundant organic osmolyte (manni<strong>to</strong>l) making<br />

only a small contribution (Trail et al., 2005).<br />

Higher turgor pressures have been recorded<br />

when asci are mounted in water (Ingold, 1939,<br />

1966).<br />

In cup fungi (discomycetes), as the asci<br />

mature they elongate and project above the


DEVELOPMENT OF ASCI<br />

243<br />

Fig 8.14 Ascospore puffing in<br />

Aleuria aurantia. Athickwhitecloud<br />

of ascospores has been released by<br />

a cluster of apothecia. Reprinted<br />

from Fuhrer (2005), with<br />

permission by Bloomings Books<br />

Pty Ltd.Original image kindly<br />

provided by B. Fuhrer.<br />

general surface of the hymenium. Their tips may<br />

be pho<strong>to</strong>tropic, as in the coprophilous fungus<br />

Ascobolus, and this ensures that the ascospores<br />

are directed upwards, <strong>to</strong>wards the light. In some<br />

discomycetes, especially those with operculate<br />

asci (e.g. Ascobolus, Peziza), large numbers of<br />

ripe asci may discharge their spores simultaneously,<br />

a phenomenon known as puffing<br />

(Fig. 8.14). This may also occur, but less obviously,<br />

in forms with inoperculate asci, e.g. Sclerotinia<br />

and Rhytisma. Puffing results in a cloud of ascospores<br />

being discharged for greater distances than<br />

with spores discharged from a single ascus<br />

(Buller, 1934; Ingold, 1971).<br />

In the flask fungi (pyrenomycetes), such as<br />

Sordaria or Podospora, as an ascus ripens it<br />

elongates and takes up a position inside the<br />

ostiole, often gripped in position by a lining layer<br />

of hairs, periphyses (Gr. prefix peri ¼ near,<br />

around, roundabout). In this case the asci<br />

discharge their spores in turn. The necks of the<br />

perithecia in Sordaria and Podospora are pho<strong>to</strong>tropic<br />

so that the ascospores are shot <strong>to</strong>wards the<br />

light. A variant of this method of discharge is<br />

found in fungi whose perithecia have necks very<br />

much longer than the length of the asci, such<br />

as Cera<strong>to</strong>s<strong>to</strong>mella and Gnomonia. Here, before<br />

discharge, the asci break at their bases and<br />

detached asci move up a canal inside the<br />

perithecial neck and are held in place within<br />

the ostiole by periphyses as they discharge their<br />

spores (Ingold, 1971).<br />

The behaviour of the bitunicate type of ascus<br />

during discharge has been described as the<br />

Jack-in-the-box mechanism (Ingold, 1971). The<br />

outer wall is relatively rigid and inextensible.<br />

As the ascus expands, the outer wall ruptures<br />

laterally or apically (see Figs. 8.12h,i) and the<br />

inner wall then stretches before the ascus<br />

explodes. Ascus discharge is thus a two-stage<br />

process. This type of mechanism is found in<br />

Loculascomycetes such as Sporormiella (Fig. 17.18),<br />

Lep<strong>to</strong>sphaeria (Fig. 17.3) and Pleospora (Figs. 17.1,<br />

17.9).<br />

In Cochliobolus the ascus is bitunicate but the<br />

endotunica is incomplete at its base, i.e. vestigially<br />

or partially bitunicate. In C. cymbopogonis the<br />

eight ascospores are spirally coiled around each<br />

other in the ascus and each has a recurved tip<br />

(Figs. 8.15a,b). The ec<strong>to</strong>ascus bursts open near<br />

its tip and the sheaf of ascospores is expelled,<br />

the incomplete endoascus forming a thimblelike<br />

cap over the tips of the ascospores as they<br />

pass through the long pseudothecial neck. The<br />

spores are not explosively discharged but<br />

are extruded, en masse, from the neck of the<br />

pseudothecium in a long tendril from which<br />

they are dispersed by rain splash. In water the<br />

spores separate from each other and push away<br />

the endotunica which earlier capped their tips<br />

<strong>to</strong>gether (see Fig. 8.15d; El-Shafie & Webster,<br />

1980; Alcorn, 1981).<br />

It is likely that in most cases the spores are<br />

spatially separated from each other as they are


244 ASCOMYCOTA (ASCOMYCETES)<br />

Fig 8.15 Ascospore liberation in Cochliobolus cymbopogonis. (a) Ascospore with a recurved tip. (b) Ascus containing a sheaf of<br />

eight spirally coiled ascospores.The ascus is bitunicate but the endoascus is not shown. (c) Ascus during the first stage of discharge.<br />

The ec<strong>to</strong>ascus has broken (arrow), the endoascus has broken at its base and has extended, remaining as a thimble-like cap over<br />

the sheaf of ascospores. (d) An ascus after the release of the ascospores which have straightened out and pushed the broken<br />

endoascus aside. (e) Diagrammatic representation of a section through a pseudothecium showing stages in ascospore release.<br />

Scale bar, (a,b) ¼ 50 mm; (c,d) ¼ 200 mm. After El-Shafie and Webster (1980).<br />

constricted on passing through the ascus pore.<br />

This has been neatly demonstrated by spinning<br />

a transparent disc over the surface of a culture<br />

of Sordaria discharging spores (Ingold &<br />

Hadland, 1959). The ascus contents are laid out<br />

on the disc in the order in which they are<br />

released. Various patterns of spore clumping and<br />

separation are visible, and although in many asci<br />

the eight spores are well separated from each<br />

other, in others there is a tendency for spores <strong>to</strong><br />

stick <strong>to</strong>gether. Calculations made from<br />

measurements of the length of the ascospore<br />

deposit and the speed of rotation of the disc, as<br />

well as by other methods, have revealed ascospores<br />

<strong>to</strong> be the fastest-accelerating biological<br />

objects (Trail et al., 2005; Vogel, 2005). The actual<br />

time taken for ascus discharge was estimated by<br />

the rotating disc method <strong>to</strong> be 0.000024 s (Ingold<br />

& Hadland, 1959). When ascospores stick<br />

<strong>to</strong>gether, they are discharged further than<br />

single-spored projectiles. In many coprophilous<br />

ascomycetes (e.g. Ascobolus, Saccobolus, Podospora)


TYPESOFFRUITBODY<br />

245<br />

the spores may be attached <strong>to</strong>gether by mucilaginous<br />

secretions and may be projected for<br />

distances of 30 cm in Ascobolus immersus and<br />

50 cm in Podospora fimicola. The distances <strong>to</strong><br />

which individual ascospores are discharged<br />

vary, but are often in the range of 1 2 cm.<br />

In some ascomycetes the ascospores are<br />

not discharged violently, and in such cases the<br />

asci are often globose instead of cylindrical.<br />

The Hemiascomycetes (Chapter 10),<br />

Plec<strong>to</strong>mycetes (Chapter 11) and several other<br />

groups have asci of this type. In Ophios<strong>to</strong>ma<br />

(Fig. 12.36) and Sphaeronaemella fimicola<br />

(Fig. 12.42) the ascus walls dissolve <strong>to</strong> release<br />

a mass of sticky spores which ooze out as a drop<br />

held in place by a ring of hairs surrounding the<br />

ostiole at the tip of a cylindrical neck which<br />

surmounts the perithecium. They are dispersed<br />

by insects. Breakdown of asci within the fruit<br />

body is also found in Chae<strong>to</strong>mium (Fig. 12.9).<br />

Ripe ascospores are extruded from the neck of<br />

the perithecium in a tendril. Possibly they are<br />

dispersed by jerking movements generated as<br />

the rough-walled perithecial hairs twist around<br />

each other. Tendrils of ascospores are sometimes<br />

found in ascomycetes which normally discharge<br />

their spores violently, e.g. Daldinia concentrica<br />

(Plate 5a), Hypocrea pulvinata (Plate 5c) and Nectria.<br />

In many marine ascomycetes the ascus walls<br />

are evanescent and dissolve <strong>to</strong> release the<br />

ascospores passively. In ascomycetes with subterranean<br />

fruit bodies, e.g. in the truffle Tuber<br />

and its relatives (Figs. 14.7, 14.8), the ascospores<br />

are not discharged violently, but are dispersed<br />

when the fruit bodies are eaten by rodents<br />

and other animals attracted by their characteristic<br />

odour.<br />

8.7 Types of fruit body<br />

The main types of ascomycete fruit body have<br />

been listed earlier (p. 21) and are drawn in<br />

Fig. 8.16). In yeasts and related fungi the asci are<br />

not enclosed by hyphae, but in most ascomycetes<br />

they are surrounded by hyphae <strong>to</strong> form an<br />

ascocarp (i.e. an ascus fruit body) or ascoma.<br />

An old term for ascus is theca (Gr. theca ¼ a case),<br />

and although this word is not now in general<br />

use, it is still found as a suffix in terms for<br />

different types of ascocarp. Byssochlamys forms<br />

clusters of naked asci (Fig. 11.15). In Gymnoascus<br />

there is a loose open network of peridial hyphae<br />

Fig 8.16 Different types of<br />

ascocarp, diagrammatic and not <strong>to</strong><br />

scale. (a) Gymnothecium made up of<br />

branched hyphae which do not<br />

completely enclose the asci.<br />

(b) Cleis<strong>to</strong>thecium completely<br />

enclosing the asci which are formed<br />

throughout the ascocarp.There is no<br />

opening. (c) Apothecium, an open cup<br />

lined by a layer of asci and associated<br />

structures forming the hymenium.<br />

(d) Perithecium with a layer of asci at<br />

the base. It opens by a pore or ostiole.<br />

Its wall or peridium is made up of<br />

flattened cells. (e) Pseudothecium.<br />

The asci are formed within locules in<br />

apseudoparenchyma<strong>to</strong>usascostroma.<br />

There is no peridium. (d) after Ingold<br />

(1971), (e) after Luttrell (1981).


246 ASCOMYCOTA (ASCOMYCETES)<br />

forming a gymnothecium (Gr. gymnos ¼ naked)<br />

and the asci can be seen through the network<br />

(Fig. 11.8). Gymnothecia are also seen in<br />

Myxotrichum (Fig. 11.10) and Ctenomyces (Fig. 11.5)<br />

where certain peridial hyphae extend as hooked<br />

hairs. In most species of Aspergillus and Penicillium<br />

which possess ascocarps, the asci are enclosed<br />

in a globose fructification with no special<br />

opening <strong>to</strong> the outside. Such ascocarps are<br />

termed cleis<strong>to</strong>carps or cleis<strong>to</strong>thecia (Gr.<br />

kleis<strong>to</strong>s ¼ enclosed). A modified cleis<strong>to</strong>thecium<br />

capable of cracking open along a line of weakness<br />

is found in the Erysiphales (powdery<br />

mildews), and this is called a chasmothecium<br />

(Gr. chasma ¼ an open mouth). In the cup fungi<br />

(Pezizales and Helotiales) as well as in many<br />

lichenized ascomycetes the asci are borne in<br />

open saucer-shaped ascocarps, and at maturity<br />

the tips of the asci are freely exposed (see Plates 6<br />

and 7). Such fruit bodies are termed apothecia (Gr.<br />

apo- ¼ away from, separate). The Pyrenomycetes<br />

(e.g. Sphaeriales and Hypocreales) have perithecia<br />

(Gr. peri- ¼ around) which are flask-shaped<br />

fruit bodies opening by a pore or ostiole (see<br />

Fig. 12.1, Sordaria fimicola). The perithecial wall<br />

is formed from sterile cells derived from hyphae<br />

which surrounded the ascogonium during development.<br />

Perithecia are often single, as in Sordaria<br />

and Neurospora, but in some genera they are<br />

embedded in or seated on a mass of tissue<br />

forming a perithecial stroma (for examples,<br />

see Plate 5). The development of pseudothecia<br />

differs from that of perithecia in that the asci<br />

are contained in one or several cavities (locules)<br />

formed within a pre-existing ascostroma (Gr.<br />

stroma ¼ mattress, bed) (Luttrell, 1981). Examples<br />

are Lep<strong>to</strong>sphaeria (Fig. 17.3) and Sporormiella<br />

(Fig. 17.18). Although the structure and development<br />

of perithecia and pseudothecia are essentially<br />

different, the term perithecium is often<br />

loosely applied <strong>to</strong> both.<br />

8.8 Fossil ascomycetes<br />

Ascomycetes are an ancient group of fungi, and<br />

fossilized structures possibly representing ascocarps<br />

made up of septate, anas<strong>to</strong>mosing hyphae<br />

have been described from the Proterozoic period<br />

about 1 billion years ago (Butterfield, 2005).<br />

Lichen-like associations between fungi and<br />

cyanobacteria or algae may have existed some<br />

600 million years ago (Yuan et al., 2005). What are<br />

believed <strong>to</strong> be the remains of perithecia have<br />

been reported from beneath the epidermis of<br />

stems and rhizomes of one of the earliest known<br />

land plants, Asteroxylon, in the Rhynie chert of<br />

the Devonian period about 400 million years ago<br />

(Taylor et al., 1999, 2005). Fossil cleis<strong>to</strong>thecia containing<br />

asci and ascospores resembling those<br />

of present-day Trichocomaceae have been<br />

found in coal balls of the Carboniferous age<br />

(Stubblefield & Taylor, 1983; Stubblefield et al.,<br />

1983). Stalked ascocarps with well-preserved<br />

ascospores have been found in amber, the<br />

fossilized resin of a conifer. They have been<br />

assigned <strong>to</strong> an extant genus Chaenothecopsis<br />

(Mycocaliciaceae). Their close resemblance of<br />

present-day species which are also associated<br />

with resin indicate little evolutionary change<br />

during the past 20 million years (Rikkinen &<br />

Poinar, 2000).<br />

On morphological grounds, Barr (1983)<br />

suggested that the ances<strong>to</strong>rs of Ascomycota<br />

should be sought among the Chytridiomycota.<br />

Confirmation of this view has since been<br />

obtained by comparison of DNA sequence<br />

data. Ascomycota are also closely related <strong>to</strong><br />

Basidiomycota, each being a derived monophyletic<br />

group (Bruns et al., 1992; Berbee & Taylor,<br />

2001). Berbee and Taylor (2001) have estimated<br />

that these two groups evolved from a common<br />

ances<strong>to</strong>r about 600 million years ago, well<br />

before the development of vascular terrestrial<br />

plants.<br />

8.9 Scientific and economic<br />

significance of ascomycetes<br />

The study of ascomycetes is of considerable<br />

scientific importance. Neurospora crassa has been<br />

the subject of intensive genetical research<br />

related <strong>to</strong> its relatively simple nutrient requirements,<br />

rapid growth, its capacity <strong>to</strong> produce


CLASSIFICATION<br />

247<br />

mutants and the ease with which it can be grown<br />

and cross-mated in culture. The dissection of<br />

ascospores from its asci by micromanipulation<br />

has enabled tetrad analysis <strong>to</strong> be performed.<br />

Research on this fungus led <strong>to</strong> the important<br />

one-gene one-enzyme concept. Budding yeast<br />

(Saccharomyces cerevisiae) and a fission yeast<br />

(Schizosaccharomyces pombe) were amongst the<br />

first eukaryotes for which the entire genome<br />

was sequenced. Studies on S. cerevisiae were basic<br />

<strong>to</strong> the understanding of the biochemistry of<br />

anaerobic respiration whilst studies of S. pombe<br />

have provided key facts by which <strong>to</strong> interpret the<br />

fundamental process of cell division, which in<br />

turn has a bearing on the understanding of the<br />

apparently uncontrolled growth of cancerous<br />

cells. The economic significance of fermentation<br />

processes involving ascomycetes and their conidial<br />

relatives is immense. Examples include alcoholic<br />

fermentations by yeasts as the basis of the<br />

wine and brewing industries, antibacterial antibiotics<br />

such as penicillin from Penicillium chrysogenum<br />

and cephalosporin from Acremonium spp.,<br />

and organic acids such as citric acid from<br />

Aspergillus niger. The immunosuppressant drug<br />

cyclosporin, which reduces the tissue rejection<br />

response and thus facilitates organ transplants,<br />

is a metabolite of Tolypocladium inflatum. Some<br />

ascomycetes are important in food production as<br />

in bread-making by yeast, cheese ripening by<br />

Penicillium roqueforti and P. camemberti and<br />

the fermentation of soybeans and wheat<br />

by Aspergillus, yeasts and bacteria <strong>to</strong> produce<br />

soy sauce. The mycoprotein Quorn is produced<br />

from mycelial biomass of Fusarium venenatum.<br />

Examples of the direct use of ascocarps as food or<br />

food flavourings are morels (Morchella spp.) and<br />

truffles (Tuber spp.).<br />

However, food spoilage may result from<br />

ascomycete contamination. A well-known example<br />

is contamination of cereal grains and grass<br />

by sclerotia of the ergot fungus Claviceps<br />

purpurea, which can cause severe, sometimes<br />

fatal, neurological, muscular and circula<strong>to</strong>ry<br />

diseases such as gangrene or abortion in cattle<br />

and man. Studies on the alkaloid <strong>to</strong>xins<br />

contained in ergot sclerotia led <strong>to</strong> the discovery<br />

of drugs useful in obstetrics and the treatment<br />

of migraine, and in the identification of the<br />

hallucinogen lysergic acid. Another potentially<br />

serious myco<strong>to</strong>xin is afla<strong>to</strong>xin produced in<br />

groundnuts, cereals and other foodstuffs<br />

infected by Aspergillus flavus. Afla<strong>to</strong>xins are<br />

highly carcinogenic in poultry and mammals,<br />

including man. Other myco<strong>to</strong>xins include zearalenone<br />

from Gibberella zeae, which causes<br />

infertility in cattle and pigs, and trichothecenes<br />

from Trichothecium roseum and Fusarium spp.,<br />

which cause aleukia in farm animals and<br />

man. A family of plant growth hormones,<br />

the gibberellins, now produced commercially,<br />

were discovered in an investigation of Bakanae<br />

(foolish seedling) disease of rice.<br />

It is not surprising that such a large group<br />

as the Ascomycota should contain numerous<br />

pathogens of plants and animals. Lifestyles<br />

are similarly varied, including biotrophic, hemibiotrophic<br />

and necrotrophic associations. Many<br />

ascomycote pathogens are of considerable economic<br />

importance.<br />

8.10 Classification<br />

It is impractical <strong>to</strong> attempt a detailed classification<br />

of ascomycetes which could include<br />

around 55 orders and 291 families (Kirk<br />

et al., 2001). We shall adopt the simplified<br />

classification outlined by M. E. Barr (2001) and<br />

Kurtzman and Sugiyama (2001). Based on a<br />

wealth of microscopic data, and especially the<br />

results of several phylogenetic analyses, five<br />

major groups (classes) of Ascomycota have<br />

been proposed, namely Archiascomycetes<br />

(Chapter 9), Hemiascomycetes (Chapter 10),<br />

Plec<strong>to</strong>mycetes (Chapter 11), Hymenoascomycetes<br />

(Chapters 12 16) and Loculoascomycetes<br />

(Chapter 17). The latter three are sometimes<br />

called ‘higher ascomycetes’ or Euascomycetes.<br />

The class Hymenoascomycetes contains ascomycetes<br />

producing asci in a hymenium, i.e. in a<br />

fertile layer around which the ascocarp develops.<br />

This is in contrast <strong>to</strong> the Loculoascomycetes,<br />

where the asci develop in a pre-formed stroma.<br />

Since the Hymenoascomycetes are a very large<br />

and diverse group, we have subdivided them in<br />

this book.


248 ASCOMYCOTA (ASCOMYCETES)<br />

Fig 8.17 Outline of a possible ascomycete phylogeny,<br />

presented as a consensus tree based on sequences of the RNA<br />

polymerase II gene. Lichenized fungi are printed in bold.<br />

Redrawn from Liu and Hall (2004), with permission. ß 2004<br />

National Academy of Sciences,U.S.A.


CLASSIFICATION<br />

249<br />

An outline of current phylogenetic relationships<br />

among the Ascomycota has been given by<br />

Liu and Hall (2004) and is shown in Fig. 8.17.<br />

More detailed delimitations of the individual<br />

groups will be discussed in subsequent chapters.<br />

In our treatment of the ascomycetes, we have<br />

attempted <strong>to</strong> consider purely asexual forms in<br />

the taxonomic context of their ascomycete state<br />

wherever practical and where known with<br />

certainty.


9<br />

Archiascomycetes<br />

9.1 <strong>Introduction</strong><br />

Several independent phylogenetic analyses of<br />

DNA sequence data (e.g. Berbee & Taylor, 1993;<br />

Sjamsuridzal et al., 1997; Liu & Hall, 2004) have<br />

grouped <strong>to</strong>gether a range of seemingly very<br />

diverse genera of ascomycetes. This group is considered<br />

<strong>to</strong> be the oldest of three broad evolutionary<br />

lineages of Ascomycota and has thus been<br />

named Archiascomycetes (Nishida & Sugiyama,<br />

1994). The core of the Archiascomycetes consists<br />

of the genera Taphrina and Pro<strong>to</strong>myces, which<br />

are facultative biotrophic plant pathogens, and<br />

the saprotrophic fission yeast Schizosaccharomyces.<br />

Also now included are the yeast-like Pneumocystis,<br />

which causes pneumonia in immunocompromised<br />

patients (see p. 259); the filamen<strong>to</strong>us<br />

fungus Neolecta, which parasitizes the roots<br />

of higher plants (Redhead, 1977; Landvik<br />

et al., 2003); and the anamorphic yeast Sai<strong>to</strong>ella.<br />

Yet other genera are included as possible<br />

members because even though their appropriate<br />

DNA sequences have not yet been obtained,<br />

they are known <strong>to</strong> be related <strong>to</strong> confirmed<br />

members. In <strong>to</strong>tal, the class Archiascomycetes<br />

currently contains some 150 species in<br />

10 genera.<br />

Because of their diverse morphological<br />

appearances and modes of life, it is difficult<br />

<strong>to</strong> describe common characters typical of the<br />

Archiascomycetes. With the exception of Neolecta,<br />

which produces apothecia, ascocarps are lacking<br />

and asci are produced individually by yeast<br />

cells or by conversion of hyphal tips. There are<br />

no differentiated ascogenous hyphae. Asexual<br />

reproduction is usually by simple division of<br />

vegetative yeast cells by budding or fission. Even<br />

in Neolecta, the apothecia are highly unusual in<br />

that they lack ascogenous hyphae and paraphyses,<br />

and in that the ascospores are capable of<br />

producing yeast-like conidia by budding while<br />

still within the ascus or after discharge (Redhead,<br />

1977). This phenomenon is also found in Taphrina<br />

(see Fig. 9.2c).<br />

The presence of Neolecta in the most basal<br />

group of ascomycetes indicates that the capacity<br />

<strong>to</strong> produce fruit bodies is probably an ancient<br />

trait. The inclusion of both yeasts and mycelial<br />

forms among the Archiascomycetes makes it<br />

impossible <strong>to</strong> decide the chicken-and-egg question<br />

as <strong>to</strong> which of these states is ancestral and<br />

which is derived. It is significant that all recent<br />

molecular studies have placed Taphrina within<br />

the Archiascomycetes because this genus has<br />

long been suspected <strong>to</strong> be close <strong>to</strong> the<br />

origin of both the higher ascomycetes and the<br />

basidiomycetes (Savile, 1968; Alexopoulos et al.,<br />

1996). Further, the position of Pneumocystis<br />

has been unclear until recently, oscillating<br />

between Basidiomycota (Wakefield et al., 1993)<br />

and Archiascomycetes (Sjamsuridzal et al., 1997;<br />

Kurtzman & Sugiyama, 2001), and thereby<br />

further supporting the suspected ancestral<br />

status of the organisms included among the<br />

Archiascomycetes.<br />

Two genera Taphrina and Schizosaccharomyces<br />

are of special significance <strong>to</strong> mycology<br />

and will be discussed more fully below, <strong>to</strong>gether<br />

with a brief account of Pneumocystis.


TAPHRINALES<br />

251<br />

9.2 Taphrinales<br />

The Taphrinales are ecologically biotrophic parasites<br />

mainly of flowering plants, causing a wide<br />

variety of disorders which often lead <strong>to</strong> strikingly<br />

abnormal development of the infected host<br />

tissue <strong>to</strong> form witches’ brooms, galls or leaf<br />

curls. About six genera are known of which<br />

Pro<strong>to</strong>myces (10 species) and Taphrina (95 species)<br />

are the most important. Both Pro<strong>to</strong>myces and<br />

Taphrina can be isolated from their hosts as<br />

ascospores, and these germinate in pure culture<br />

by budding <strong>to</strong> form saprotrophic haploid yeast<br />

cells. In the host plant, however, a mycelium<br />

of intercellular septate hyphae is produced. In<br />

Pro<strong>to</strong>myces, hyphae are diploid, whereas they are<br />

dikaryotic in Taphrina. Dikaryotic hyphae are<br />

most unusual among ascomycetes but are typical<br />

of basidiomycetes. Biotrophic infection of the<br />

host plant culminates in individual hyphal tips<br />

undergoing meiosis (preceded by karyogamy<br />

in Taphrina), producing usually eight haploid<br />

ascospores which are discharged violently. We<br />

will discuss only Taphrina here; for Pro<strong>to</strong>myces and<br />

related genera, species descriptions are given<br />

by Reddy and Kramer (1975).<br />

9.2.1 Taphrina<br />

Species of Taphrina are mostly parasitic on<br />

Fagaceae and Rosaceae (Mix, 1949), causing<br />

diseases of three main kinds. (1) Leaf curl or<br />

blister diseases, e.g. Taphrina deformans, the cause<br />

of peach leaf curl (Fig. 9.1; Plate 4a); T. <strong>to</strong>squinetii,<br />

the cause of leaf blister of alder; and T. populina,<br />

the cause of yellow leaf blister of poplar. (2)<br />

Diseases of above-ground plant organs in which<br />

the infected twig undergoes repeated branching<br />

<strong>to</strong> form dense tufts of twigs called witches’<br />

brooms. Examples are T. betulina, causing<br />

witches’ brooms of birch (Plate 4b), T. insititiae<br />

causes witches’ brooms of plum and damson,<br />

and T. wiesneri causes witches’ brooms and leaf<br />

curl of cherry. However, not all witches’ brooms<br />

are caused by Taphrina, and similar twig proliferation<br />

is also associated with infection by mites.<br />

(3) Diseases of fruits, e.g. T. pruni, which causes<br />

the condition known as pocket plums in which<br />

the fruit is wrinkled and shrivelled and has a<br />

cavity in the centre in place of the s<strong>to</strong>ne. Taphrina<br />

amen<strong>to</strong>rum causes conspicuous <strong>to</strong>ngue-like<br />

outgrowths on female catkins of alder, Alnus<br />

glutinosa (Plate 4c).<br />

Taphrina deformans<br />

Peach leaf curl is common on leaves and twigs<br />

on peach and almond, especially after a cool and<br />

moist spring. Towards the end of May, infected<br />

peach leaves show raised reddish puckered<br />

blisters which eventually acquire a waxy bloom<br />

(Fig. 9.1). Sections of leaves in this condition<br />

show an extensive septate mycelium growing<br />

between cells of the mesophyll and between the<br />

cuticle and epidermis, where the hyphae end in<br />

Fig 9.1 Taphrina deformans. Peach leaf<br />

showing leaf curl.


252 ARCHIASCOMYCETES<br />

swollen tips which have been termed chlamydospores<br />

by Martin (1940) but are, in fact, ascus<br />

initials or ascogenous cells (see Fig. 9.2). The<br />

interface between the parasitic fungus and the<br />

host takes the form of contact between their<br />

walls. No specialized haus<strong>to</strong>ria have been found<br />

in T. deformans (Syrop, 1975), although they have<br />

been reported in some other species. Cy<strong>to</strong>logical<br />

studies (Martin, 1940; Kramer, 1961; Syrop &<br />

Beckett, 1976) have revealed that the segments of<br />

the mycelium and the young ascogenous cells<br />

are mostly binucleate. If the cells are multinucleate,<br />

then the nuclei are at least arranged in<br />

pairs (Syrop & Beckett, 1976). In the ascogenous<br />

cell, the two nuclei fuse and the diploid nucleus<br />

divides mi<strong>to</strong>tically. The upper of the two daughter<br />

nuclei then undergoes meiosis followed by a<br />

mi<strong>to</strong>sis so that eight nuclei result, which form<br />

the nuclei of the eight ascospores. The lower<br />

daughter nucleus remains in the lower part of<br />

the ascogenous cell and is often separated from<br />

the upper nucleus by a cross wall. During these<br />

nuclear divisions, the wall of the ascogenous<br />

cell has stretched <strong>to</strong> form an ascus. Delimitation<br />

of the ascospores occurs at the eight-nucleate<br />

stage. The individual nuclei become enclosed<br />

by double-delimiting membranes which do not<br />

arise from the nuclear envelope as in most<br />

ascomycetes, but by invagination of the plasmalemma<br />

of the developing ascus (Syrop & Beckett,<br />

1972). Within the ascus, the ascospores may<br />

bud so that ripe asci may contain numerous<br />

yeast cells (see Fig. 9.2c). These yeast cells can be<br />

regarded as the anamorphic state of Taphrina,<br />

and they have been named Lalaria (Moore, 1990;<br />

Inácio et al., 2004). The asci form a palisade-like<br />

layer above the epidermis, and it is their presence<br />

which gives the infected leaf its waxy<br />

bloom.<br />

The ascospores and yeast cells are projected<br />

from the ascus which often opens by a characteristic<br />

slit (Fig. 9.2c). Yarwood (1941) has<br />

shown that there is a diurnal cycle of ascus<br />

development and discharge in T. deformans.<br />

Nuclear fusion takes place during the afternoon<br />

or evening; nuclear divisions are complete by<br />

Fig 9.2 Taphrina deformans. (a) T.S. peach<br />

leaf showing intercellular mycelium and<br />

subcuticular ascogenous cells. (b) T.S. peach<br />

leaf showing ascogenous cells and asci,<br />

containing eight ascospores. (c) T.S. leaf<br />

showing a dehisced ascus, an eight-spored<br />

ascus and an ascus in which the ascospores<br />

are budding. Ascospores budding outside<br />

theascusarealsoshown.(d j) Cy<strong>to</strong>logy of<br />

ascus formation (after Martin,1940). (d,e).<br />

Fusion of nuclei in ascogenous cell.<br />

(f) Elongating ascogenous cell containing<br />

two nuclei formed by mi<strong>to</strong>sis from the<br />

fusion nucleus.The upper nucleus has begun<br />

<strong>to</strong> divide meiotically. (g) Uninucleate ascus<br />

with uninucleate basal cell. (h,i) Four- and<br />

eight-nucleate asci. (j) Binucleate germ tube<br />

in germinating ascospore.


SCHIZOSACCHAROMYCETALES<br />

253<br />

about 5 a.m. and the spores appear mature by<br />

8 a.m. However, maximum spore discharge does<br />

not occur until about 8 p.m. Outside the ascus,<br />

ascospores or yeast cells may continue budding<br />

and the fungus can be grown saprotrophically<br />

as a yeast in agar or liquid culture. The yeast<br />

cells are often pigmented due <strong>to</strong> the presence of<br />

b-carotene and several other carotenoid pigments<br />

(van Eijik & Roeymans, 1982). Young<br />

leaves can be infected from such yeast cells and<br />

it has been shown that a culture derived from<br />

a single ascospore can cause infection resulting<br />

in the formation of a fresh crop of asci, so that<br />

T. deformans is homothallic. In this respect it differs<br />

from some other species, e.g. T. epiphylla,<br />

where the fusion of yeast cells, presumably of<br />

different mating types, is necessary before infection<br />

can occur (Kramer, 1987). In T. deformans the<br />

binucleate condition is established at the first<br />

nuclear division of a yeast cell placed on a peach<br />

leaf, and the two daughter nuclei remain associated<br />

in the germ tube which penetrates the<br />

cuticle (Fig. 9.2j). In other Taphrina species, the<br />

germ tube penetrates through s<strong>to</strong>mata but is<br />

unable <strong>to</strong> breach the intact cuticle (Taylor &<br />

Birdwell, 2000).<br />

9.2.2 Growth hormones<br />

In infections of peach leaves with T. deformans,<br />

the dis<strong>to</strong>rtions of the host tissue are associated<br />

with division and hypertrophy of the cells of the<br />

palisade mesophyll. In liquid cultures, especially<br />

on media containing tryp<strong>to</strong>phane, considerable<br />

quantities of the auxin-type phy<strong>to</strong>hormone<br />

indole acetic acid (IAA) have been demonstrated.<br />

A number of different cy<strong>to</strong>kinins are also<br />

produced by several species of Taphrina in culture<br />

(Kern & Naef-Roth, 1975; Tudzynski, 1997).<br />

Together, these hormones promote processes of<br />

cell division, enlargement and differentiation in<br />

plants, and leaves infected with T. deformans<br />

show higher levels of auxins and cy<strong>to</strong>kinins<br />

than uninfected leaves (Sziráki et al., 1975). It is<br />

therefore tempting <strong>to</strong> assume that the fungus<br />

produces these substances also in planta.<br />

However, this has not been formally proven<br />

yet, and the Taphrina peach system seems <strong>to</strong><br />

have been less thoroughly examined than the<br />

interaction between Plasmodiophora brassicae and<br />

cabbage plants (see p. 63).<br />

9.2.3 Control of Taphrina deformans<br />

Taphrina deformans is by far the most serious<br />

pathogen among the Taphrinales, and it occurs<br />

wherever peach or almond trees grow. It is not<br />

yet entirely clear how T. deformans overwinters;<br />

Butler and Jones (1949) and Smith et al. (1988)<br />

considered it unlikely that the mycelial form is<br />

involved because leaves harbouring mycelium<br />

are shed in the autumn. It is more probable<br />

that yeast cells arising from discharged ascospores<br />

survive saprotrophically on the surface of<br />

twigs or in bud scales. Between November and<br />

March, the yeast cells develop thick walls and in<br />

spring, as the peach buds open, they produce<br />

germ tubes which penetrate the young leaves.<br />

The first symp<strong>to</strong>ms of infection can be seen as<br />

soon as the buds break, but no further infection<br />

occurs from about early July onwards. This may<br />

be because T. deformans has a relatively low<br />

temperature maximum of 26 30°C (Butler &<br />

Jones, 1949).<br />

Good chemical control of T. deformans can<br />

be achieved by spraying with Bordeaux mixture<br />

in autumn after leaf fall, in order <strong>to</strong> reduce the<br />

population of yeast cells on the twigs. Another<br />

spray in early spring, at the time of bud swelling,<br />

will give improved control of infection because<br />

the time span in which Taphrina can infect<br />

is limited. Dithiocarbamates or other simple<br />

fungicides are commonly used in spring<br />

(Smith et al., 1988). Other diseases caused by<br />

Taphrina can be controlled in a similar way if<br />

necessary.<br />

9.3 Schizosaccharomycetales<br />

The classification of the Schizosaccharomycetales<br />

has been the subject of controversial<br />

discussions, but the emerging consensus is<br />

that there is only one genus with three species,<br />

S. japonicus, S. oc<strong>to</strong>sporus and S. pombe (Kurtzman<br />

& Robnett, 1998; Vaughan-Martini & Martini,<br />

1998a; Barnett et al., 2000). All three species grow


254 ARCHIASCOMYCETES<br />

as saprotrophic yeasts which reproduce asexually<br />

by fission, i.e. by division of a vegetative cell<br />

in<strong>to</strong> two daughter cells of equal size (Fig. 9.3a).<br />

Schizosaccharomyces is therefore called the fission<br />

yeast. Occasionally, especially in S. japonicus, true<br />

septate hyphae can be formed, and these may<br />

fragment in<strong>to</strong> arthrospores. Sexual reproduction<br />

is by conjugation of two haploid vegetative<br />

yeast cells, followed by karyogamy and meiosis<br />

which gives rise <strong>to</strong> four or, more usually, eight<br />

ascospores (Figs. 9.3a,b).<br />

Schizosaccharomyces can be isolated from substrates<br />

rich in soluble carbon sources, e.g. tree<br />

exudates, fruits, honey and fruit juices. The bestknown<br />

species are S. oc<strong>to</strong>sporus and S. pombe. The<br />

latter is the fermenting agent of African millet<br />

beer (pombe) and arak in Java. It can <strong>to</strong>lerate<br />

ethanol levels up <strong>to</strong> 7% (v/v). Both species grow<br />

well in liquid culture or on solid media such as<br />

malt extract agar, developing ripe asci within<br />

3 days at 25°C. All stages of the life cycle can be<br />

readily seen if a preparation of cells from an agar<br />

culture of S. oc<strong>to</strong>sporus is made in water (Fig. 9.3).<br />

Individual cells are globose <strong>to</strong> cylindrical,<br />

uninucleate and haploid. Cell division is<br />

preceded by intranuclear mi<strong>to</strong>sis, <strong>to</strong>wards the<br />

end of which the nucleus constricts and becomes<br />

dumb-bell shaped (Tanaka & Kanbe, 1986). The<br />

division of the cell in<strong>to</strong> two daughter cells is<br />

brought about by the centripetal development of<br />

a septum which cuts the cy<strong>to</strong>plasm in<strong>to</strong> two. The<br />

two sister cells may remain attached <strong>to</strong> each<br />

other for a while, or may separate by breakdown<br />

of a layer of material in the middle of the septum<br />

(Sipiczki & Bozsik, 2000).<br />

Ascus formation in S. pombe is preceded by<br />

copulation. Schizosaccharomyces oc<strong>to</strong>sporus is homothallic,<br />

and quite often adjacent sister cells<br />

may fuse <strong>to</strong>gether. In the case of S. pombe, both<br />

homothallic and heterothallic strains are known,<br />

the latter with a bipolar mating system (hþ and<br />

h mating types). When cells of opposite mating<br />

type of S. pombe are grown <strong>to</strong>gether in liquid<br />

culture, especially under conditions of nitrogen<br />

starvation, a strong sexual agglutination<br />

occurs. This clumping <strong>to</strong>gether of the cells<br />

becomes visible as a flocculation of the culture.<br />

Changes in cell surface properties are<br />

Fig 9.3 Schizosaccharomyces oc<strong>to</strong>sporus.<br />

(a) Vegetative cells, three of which showing<br />

transverse division.Two cells <strong>to</strong> the right of<br />

the picture are conjugating. (b) Four- and<br />

eight-spored asci.


SCHIZOSACCHAROMYCETALES<br />

255<br />

important in agglutination, and fimbriae have<br />

been observed at the surface of cells stimulated<br />

by the appropriate pheromone (Johnson et al.,<br />

1989). Using stable heterothallic haploid strains,<br />

the purification of the pheromones of S. pombe<br />

has been achieved; both are linear oligopeptide<br />

hormones (Davey, 1992; Imai & Yamamo<strong>to</strong>, 1994).<br />

The binding of a hormone released by cells of one<br />

mating type <strong>to</strong> recep<strong>to</strong>rs in the membranes of<br />

cells of opposite mating type initiates a signalling<br />

chain which in turn triggers the cellular<br />

response leading <strong>to</strong> agglutination and conjugation<br />

(Davey, 1998). The principle of mating<br />

fac<strong>to</strong>rs will be discussed in more detail for<br />

Saccharomyces cerevisiae (p. 266).<br />

During agglutination, two cells come in<strong>to</strong><br />

contact by a portion of their cell wall. In homothallic<br />

strains, the fusing cells are often sister<br />

cells formed by preceding mi<strong>to</strong>tic division.<br />

A pore is formed in the centre of the attachment<br />

area and this widens and elongates <strong>to</strong> form a<br />

conjugation canal (Fig. 9.3a). During this process<br />

the nuclei, one from each cell, migrate <strong>to</strong>wards<br />

each other and fuse. Vacuoles may appear in the<br />

young ascus following nuclear fusion. The fused<br />

nucleus elongates and may reach half the length<br />

of the ascus, and then divides by constriction,<br />

the nuclear membrane remaining intact during<br />

division. The two daughter nuclei migrate <strong>to</strong><br />

opposite ends of the ascus and divide further.<br />

These two divisions constitute meiosis. A single<br />

mi<strong>to</strong>tic division usually follows so that eight<br />

haploid nuclei result, and eight ascospores are<br />

finally differentiated (Tanaka & Hirata, 1982).<br />

Four-spored asci are also common. The ascospores<br />

are released passively following disintegration<br />

of the ascus wall.<br />

The life cycle of Schizosaccharomyces (Fig. 9.4) is<br />

thus interpreted as being based on haploid<br />

vegetative cells which fuse <strong>to</strong> form asci, the<br />

only diploid cells. Meiosis in the ascus res<strong>to</strong>res<br />

the haploid condition. Some variation in this<br />

pattern may occur. For example, in S. japonicus<br />

and S. pombe, limited division of the zygote in the<br />

diploid state before ascospore formation may<br />

take place, and it is possible <strong>to</strong> select diploid<br />

strains (Tange & Niwa, 1995).<br />

Physical and chemical analyses of the<br />

cell walls of Schizosaccharomyces show that they<br />

are principally composed of a b-(1,3)-glucan<br />

with b-(1,6)-branches making up 50 54% of<br />

the <strong>to</strong>tal cell wall carbohydrates, and of an<br />

a-(1,3)-glucan with a-(1,4)-branches contributing<br />

28 32%. There are also trace amounts of a<br />

branched b-(1,6)-glucan (Manners & Meyer, 1977;<br />

Kopecka et al., 1995). The b-(1,3)-glucan synthase is<br />

involved in all aspects of wall synthesis, including<br />

polarized growth, septum formation, and the<br />

formation and germination of ascospores (Cortés<br />

et al., 2002). A galac<strong>to</strong>mannan linked <strong>to</strong> wall<br />

matrix glycoprotein makes up about 9 14% of<br />

the cell wall polysaccharides (Manners & Meyer,<br />

1977). As in other Archiascomycetes, chitin is<br />

present only in traces (Sietsma & Wessels, 1990),<br />

but it seems <strong>to</strong> play an important role in<br />

ascospore formation (Arellano et al., 2000). The<br />

walls of ascospores contain amylose and give a<br />

blue reaction with iodine.<br />

Fig 9.4 The life cycle of the<br />

homothallic yeast<br />

Schizosaccharomyces oc<strong>to</strong>sporus.<br />

Small open circles represent<br />

haploid nuclei; diploid nuclei are<br />

larger and split. Key events in the<br />

life cycle are plasmogamy (P),<br />

karyogamy (K) and meiosis (M).


256 ARCHIASCOMYCETES<br />

In the following sections we give brief summaries<br />

of areas in which research on S. pombe is<br />

of outstanding significance for the discipline<br />

of biology as a whole. We anticipate that the<br />

relevance of this yeast for fundamental research<br />

will further increase in the future. Whilst we do<br />

not believe in the concept of a ‘model organism’<br />

or even a ‘model fungus’, it is becoming<br />

clear that S. pombe, based on its more ancestral<br />

position in the phylogenetic system, is in many<br />

ways more relevant <strong>to</strong> the study of eukaryotic<br />

biology than its great rival, the more derived<br />

Saccharomyces cerevisiae (see p. 270). The entire<br />

genomes of both yeasts have been sequenced,<br />

and research is under way with S. pombe <strong>to</strong> find<br />

out the minimum number of genes (approximately<br />

17.5% of all genes) required for the<br />

basic functioning of this organism (Decottignies<br />

et al., 2003).<br />

9.3.1 Schizosaccharomyces pombe and<br />

the cell cycle<br />

The term ‘cell cycle’ denotes a carefully controlled<br />

sequence of regula<strong>to</strong>ry and biosynthetic<br />

processes which guide a cell arising from mi<strong>to</strong>sis<br />

<strong>to</strong>wards its division in<strong>to</strong> two daughter cells.<br />

Research on S. pombe has given us a fundamental<br />

understanding of the cell cycle. The literature<br />

on this <strong>to</strong>pic is vast, and it is beyond the scope<br />

of this book <strong>to</strong> give more than the briefest of<br />

summaries. Our account borrows heavily from<br />

the textbook by Lewin (2000), which also provided<br />

the basis of the diagrammatic summary<br />

(Fig. 9.5).<br />

A young cell arising from mi<strong>to</strong>tic division<br />

starts its life in the G1 phase (G ¼ gap) and may<br />

synthesize RNA, protein and other cellular constituents.<br />

It may grow in size but it does not<br />

duplicate its DNA at this point. The first crucial<br />

control point of the cell cycle is the<br />

START point, located in G1. At this point, the<br />

cell becomes committed <strong>to</strong> mi<strong>to</strong>sis, and other<br />

options notably sexual reproduction are no<br />

longer available, i.e. beyond the START point the<br />

cell becomes insensitive <strong>to</strong> mating pheromones.<br />

When DNA duplication is actually initiated,<br />

the cell moves from the G1 <strong>to</strong> the S (synthesis<br />

of DNA) phase. After DNA replication has been<br />

completed, the G2 phase follows, during<br />

which the S. pombe cell further enlarges in size<br />

and produces all organelles and macromolecules<br />

which are required <strong>to</strong> support two daughter<br />

cells. A second control point is the boundary<br />

between G2 and the M (mi<strong>to</strong>tic) phase; when<br />

this has been passed, the cell s<strong>to</strong>ps elongating.<br />

Condensation and separation of the chromosomes<br />

occur, followed by septation and physical<br />

separation of the two daughter cells. The identification<br />

of genes whose products are involved<br />

in the regulation of the cell cycle was possible<br />

by analysing temperature-sensitive mutants,<br />

i.e. mutants which grow normally at reduced<br />

(permissive) temperature (e.g. 25°C) but are<br />

blocked at some stage of the cell cycle at a<br />

higher (restrictive) temperature (e.g. 37°C).<br />

The most fundamental gene involved in the<br />

cell cycle is cdc2 because its product a protein<br />

kinase is involved at both the START and G2/M<br />

control points, and it is now known <strong>to</strong> fulfil<br />

the same universal role in all eukaryotes, including<br />

humans (Lee & Nurse, 1987; Nurse, 1990).<br />

In order <strong>to</strong> act in such a way, the cdc2 protein<br />

(written as Cdc2) combines with different<br />

proteins at specific stages of the cell cycle.<br />

These proteins are termed cyclins because their<br />

levels in the yeast cell show one peak in each cell<br />

cycle, followed by their degradation or inactivity.<br />

There are G1 cyclins and G2 cyclins which have<br />

different properties in combination with Cdc2.<br />

However, the activity of Cdc2 is modulated not<br />

only by the binding of cyclins, but also by kinases<br />

or phosphatases which, respectively, phosphorylate<br />

or dephosphorylate the Cdc2 protein. These<br />

respond <strong>to</strong> environmental stimuli and often<br />

antagonize each other in their effects on Cdc2.<br />

This allows a fine-tuning of the cell cycle in<br />

response <strong>to</strong> environmental fac<strong>to</strong>rs such as the<br />

presence of pheromones which would prevent<br />

progression through START, or nutrient availability.<br />

Whilst the regulation of Cdc2 is relatively<br />

well unders<strong>to</strong>od, few of its substrates have<br />

been identified as yet, and this is an area of<br />

ongoing research.<br />

An understanding of the cell cycle of<br />

S. pombe is of significance far beyond mycology<br />

because the principles are conserved across all<br />

eukaryotes. In mammals, one regula<strong>to</strong>ry fac<strong>to</strong>r


SCHIZOSACCHAROMYCETALES<br />

257<br />

Fig 9.5 ThecentralroleofCdc2inthecellcycleofS. pombe.The START checkpoint is passed only if Cdc2 is combined with a<br />

G1 cyclin (cig2) and is phosphorylated at a threonine residue at position161 (Thr) but dephosphorylated at a tyrosine residue at<br />

position15 (Tyr).The second major checkpoint is between the G2 and M phases; here Cdc2 needs <strong>to</strong> be coupled with a G2 cyclin<br />

(Cdc13) and must be phosphorylated at Thr but dephosphorylated at Tyr.The cell cycle is therefore controlled by the type of cyclin<br />

available for coupling with Cdc2, and by the action of kinases such as wee1or mik1and phosphatases (e.g.Cdc25). Adapted from<br />

Lewin (2000).<br />

which s<strong>to</strong>ps the cell cycle at the G2/M control<br />

point and commits the cell <strong>to</strong> apop<strong>to</strong>sis (selfdestruction)<br />

is DNA damage. This recognition<br />

mechanism is a most important protection<br />

against uncontrolled cell growth (cancer). These<br />

and other implications of the work on the cell<br />

cycle of S. pombe have resulted in the award of<br />

the Nobel Prize for Medicine and Physiology,<br />

among others, <strong>to</strong> Sir Paul Nurse in 2001. His<br />

Nobel lecture (Nurse, 2002) is a stimulating and<br />

readable account of the unravelling of the cell<br />

cycle in S. pombe.


258 ARCHIASCOMYCETES<br />

Fig 9.6 The cy<strong>to</strong>skele<strong>to</strong>n and the cell cycle in<br />

Schizosaccharomyces pombe. A new cell initially grows only at<br />

the old end (a) before bipolar growth is assumed (b). In<br />

growing cells, microtubules (dark lines) are orientated<br />

longitudinally, forming a basket around the nucleus (large<br />

sphere) and projecting in<strong>to</strong> the poles. Actin patches (white<br />

dots) are located in the growing tips, as they are in<br />

filamen<strong>to</strong>us fungi. At mi<strong>to</strong>sis (c), microtubules form the<br />

intranuclear spindle while actin aggregates <strong>to</strong> form a cortical<br />

ring in the vicinity of the dividing nucleus (d,e).The ring<br />

constricts as the wall of the septum is laid down; upon<br />

completion of nuclear migration in<strong>to</strong> the poles, the nuclear<br />

spindlebreaksdownandthebaske<strong>to</strong>fcy<strong>to</strong>plasmic<br />

microtubules reforms (f). After cell fission has been<br />

completed (g), actin relocates in<strong>to</strong> the old end (h) at which<br />

growth is resumed. Redrawn from Brunner and Nurse (2000),<br />

Philosophical Transactions ofthe Royal Society,bycopyright<br />

permission of The Royal Society.<br />

9.3.2 Morphogenesis in S. pombe<br />

Many aspects of the cell biology and ultrastructure<br />

of S. pombe have been studied extensively,<br />

and the results pertaining <strong>to</strong> the cy<strong>to</strong>skele<strong>to</strong>n<br />

are of particular relevance <strong>to</strong> filamen<strong>to</strong>us fungi.<br />

Freshly divided cells of S. pombe grow only<br />

at one end (the ‘old end’ opposite the septum),<br />

i.e. growth is initially monopolar (Fig. 9.6a)<br />

before bipolar growth starts in early G2 phase<br />

(Fig. 9.6b). In growing cells, microtubules are<br />

located in the cy<strong>to</strong>plasm and are orientated<br />

longitudinally, enclosing the nucleus like a<br />

basket. Mutant and inhibi<strong>to</strong>r studies have<br />

revealed a crucial role for microtubules in<br />

co-ordinating polarized growth, i.e. in focusing<br />

cell wall extension <strong>to</strong> either or both of the<br />

two opposite poles (Brunner & Nurse, 2000;<br />

Hayles & Nurse, 2001). Microtubules seem <strong>to</strong> be<br />

involved in transporting the Tea1 protein <strong>to</strong> the<br />

poles. This protein acts as a termination signal<br />

for microtubule elongation, and by<br />

attracting microtubules it effectively controls<br />

its own transport (Mata & Nurse, 1998; Sawin<br />

& Nurse, 1998). Interactions between Tea1p<br />

and other proteins mark the cell end, i.e. the<br />

site of Tea1p accumulation, thereby fixing the<br />

growth direction in S. pombe (Niccoli et al., 2003;<br />

Sawin & Snaith, 2004). Actin is also located in the<br />

growing poles of S. pombe. We can speculate<br />

that actin filaments and microtubules fulfil<br />

a similar role in S. pombe as in the apices of<br />

filamen<strong>to</strong>us fungi, with microtubules mediating<br />

long-distance transport and actin fine-tuning<br />

the fusion of secre<strong>to</strong>ry vesicles with the plasma<br />

membrane.<br />

When mi<strong>to</strong>sis starts, there is a complete<br />

remodelling of the cy<strong>to</strong>skele<strong>to</strong>n, with the disappearance<br />

of cy<strong>to</strong>plasmic microtubules and the


PNEUMOCYSTIS<br />

259<br />

formation of the intranuclear spindle (Hagan,<br />

1998). No further cell wall extension takes<br />

place during mi<strong>to</strong>sis, presumably because no<br />

cy<strong>to</strong>plasmic microtubules are available <strong>to</strong> drive<br />

it. By contrast, actin relocates from the poles <strong>to</strong><br />

the centre of the cell, forming a ring around<br />

the equa<strong>to</strong>r in close proximity <strong>to</strong> the nucleus<br />

(Figs. 9.6c,d). Actin aggregation is co-ordinated<br />

by a protein (Mid1p) emitted from the nucleus<br />

<strong>to</strong> form a broad cortical band, which in turn<br />

is controlled by the activity of the nuclear protein<br />

kinases Plo1p and Pdk1p (Bähler et al., 1998;<br />

Brunner & Nurse, 2000; Bimbó et al., 2005).<br />

Since the positioning of the nucleus in the<br />

cell is determined by microtubules, these are<br />

ultimately responsible for morphogenesis in<br />

S. pombe. Microtubules pull apart the two<br />

daughter nuclei (Figs. 9.6d,e), and when these<br />

have reached the two opposite poles (Fig. 9.6e),<br />

the spindle breaks down and the basket of<br />

cy<strong>to</strong>plasmic microtubules is re-established (Fig.<br />

9.6f). Meanwhile, the actin ring co-ordinates the<br />

inward growth of the septum wall. When the<br />

septum has been completed, actin relocates <strong>to</strong><br />

the two old ends, one in each daughter<br />

cell, which resume growth (Figs. 9.6g,h). It is<br />

remarkable that the septum is laid down<br />

precisely in the middle of a cell which was<br />

itself synthesized by asymmetric elongation of<br />

its two ends.<br />

If a cell of S. pombe comes in<strong>to</strong> close proximity<br />

<strong>to</strong> a cell of opposite mating type and is in the G1<br />

phase prior <strong>to</strong> the START point, it will conjugate.<br />

The formation of a projection tip during conjugation<br />

requires the presence of actin for localized<br />

cell wall synthesis and lysis, and microtubules<br />

<strong>to</strong> localize actin <strong>to</strong>wards this new if transient<br />

growing point (Petersen et al., 1998). Cell-<strong>to</strong>-cell<br />

fusion also requires the accumulation of actin<br />

(Kurahashi et al., 2002).<br />

9.4 Pneumocystis<br />

This is an unusual but appropriately named<br />

group of organisms living as cyst-like cells in<br />

the lungs of mammalian hosts, where they<br />

cause pneumonia in immunocompromised<br />

individuals. The identity of these organisms as<br />

fungi was established beyond doubt only relatively<br />

recently, following DNA sequence comparisons.<br />

Recently, these techniques have also<br />

permitted the distinction between different<br />

taxonomic entities within Pneumocystis which<br />

correlate with the taxonomy of their hosts.<br />

Thus, the original name P. carinii is now applied<br />

by most workers only <strong>to</strong> a species infecting rats,<br />

with the human pathogen called P. jirovecii<br />

(Stringer et al., 2002).<br />

Little is known about the life cycle of<br />

Pneumocystis because this organism cannot<br />

be grown satisfac<strong>to</strong>rily outside its host. Cushion<br />

(2004) summarized evidence indicating that<br />

there is probably a haploid trophic phase in<br />

which cells divide by binary fission in an amoebalike<br />

way, i.e. by constriction. Trophic cells of<br />

Pneumocystis are firmly attached <strong>to</strong> mammalian<br />

pneumocyte I cells in the alveolar regions of<br />

lung tissue. The cell wall of the trophic stage<br />

is unusually thin and flexible. If two compatible<br />

trophic cells fuse, a diploid zygote is formed<br />

and undergoes meiosis, producing a thick-walled<br />

cyst containing eight ascospores which are<br />

presumed <strong>to</strong> develop in<strong>to</strong> a fresh crop of trophic<br />

cells upon germination. Pneumocystis can<br />

cause lethal pneumonia in immunocompromised<br />

hosts and the pneumonia it causes is<br />

regarded as an AIDS-defining illness. Infections<br />

have also been linked with the sudden infant<br />

death syndrome, reflecting contact of children<br />

with Pneumocystis very soon after birth. Exposure<br />

<strong>to</strong> inoculum seems <strong>to</strong> have little effect on<br />

immunocompetent individuals, although there<br />

is evidence that they may carry latent Pneumocystis<br />

infections of limited duration. Considerable<br />

uncertainty exists about the epidemiology<br />

of Pneumocystis. Although there have been occasional<br />

reports of the detection of P. jirovecii<br />

DNA in nature, there is no convincing evidence<br />

of any external reservoir of inoculum which<br />

could represent a source of human infections.<br />

The fungus therefore seems <strong>to</strong> be spread mainly<br />

or exclusively between humans. One way by<br />

which Pneumocystis may avoid the mammalian<br />

immune response is its ability <strong>to</strong> alter the<br />

properties of surface glycoproteins acting as<br />

antigens.


260 ARCHIASCOMYCETES<br />

There are many oddities about Pneumocystis<br />

(Stringer, 1996, 2002). One is that this fungus<br />

lacks ergosterol, utilizing cholesterol instead<br />

as its major membrane sterol. This explains<br />

the insensitivity of Pneumocystis <strong>to</strong> amphotericin<br />

B, the most important drug against fungal<br />

infections of humans (see Fig. 10.9). In contrast,<br />

although the cell wall of trophic cells of<br />

Pneumocystis is unusually thin, this fungus is susceptible<br />

<strong>to</strong> inhibi<strong>to</strong>rs of b-(1,3)-glucan synthesis<br />

such as echinocandins (Schmatz et al., 1990).<br />

<strong>Third</strong>ly, there are only two rRNA gene repeat<br />

units, in contrast <strong>to</strong> other fungi which contain<br />

some 50 250 copies of it (see Fig. 1.24).


10<br />

Hemiascomycetes<br />

10.1 <strong>Introduction</strong><br />

The class Hemiascomycetes contains the classical<br />

ascomycete yeasts, exclusive of those which<br />

belong <strong>to</strong> the Archiascomycetes (see the preceding<br />

chapter) and the ‘black yeasts’ such as<br />

Aureobasidium (see p. 486). Detailed descriptions<br />

of the individual yeast genera and species are<br />

given in Kurtzman and Fell (1998) and Barnett<br />

et al. (2000). A useful taxonomic overview is that<br />

by Kurtzman and Sugiyama (2001). There is only<br />

one order, the Saccharomycetales, which has<br />

been divided in<strong>to</strong> 11 families and 276 species<br />

(Kirk et al., 2001; Kurtzman & Sugiyama, 2001).<br />

However, detailed phylogenetic analyses of the<br />

Hemiascomycetes (Kurtzman & Robnett, 1998,<br />

2003) indicate that this family arrangement is<br />

likely <strong>to</strong> be modified in the future, and for this<br />

reason we shall focus on selected genera.<br />

The key feature that distinguishes the Hemiand<br />

Archiascomycetes from the higher ascomycetes<br />

(Euascomycetes) is that ascogenous<br />

hyphae and an ascocarp, i.e. an investment of<br />

sterile hyphae surrounding the asci, are lacking<br />

in the first two groups. Instead, the asci are<br />

formed freely and singly, either directly following<br />

karyogamy or more rarely after a prolonged<br />

diploid phase. Another distinguishing feature<br />

is the composition of the cell wall, which<br />

contains very little chitin in the Hemi- and<br />

Archiascomycetes. Chitin is often confined <strong>to</strong> a<br />

small ring around the site where the daughter<br />

cell is produced (the bud scar). An ultrastructural<br />

feature of distinction concerns the septal pore<br />

of any hypha that may be produced. One or<br />

several pores may be present, and these are<br />

usually very small or plugged. They lack Woronin<br />

bodies, in contrast <strong>to</strong> Euascomycete septa<br />

which usually have only one large pore with<br />

associated Woronin bodies (Alexopoulos et al.,<br />

1996; M. E. Barr, 2001; see Fig. 8.3). Hence, the<br />

Euascomycete septal pore permits passage of<br />

organelles including nuclei (see Fig. 8.2), whereas<br />

the micropore of the Hemiascomycete septum<br />

does not. Cy<strong>to</strong>plasmic communication between<br />

adjacent hyphal cells therefore does not seem<br />

<strong>to</strong> be possible.<br />

It is impossible <strong>to</strong> give a watertight set of<br />

criteria by which Hemiascomycetes can be distinguished<br />

from Archiascomycetes. The predominant<br />

growth form of Hemiascomycetes in culture<br />

as well as in nature is the yeast state, although<br />

a limited mycelium or pseudomycelium may<br />

also be present. Archiascomycetes may grow as<br />

a mycelium in nature but as yeasts in the labora<strong>to</strong>ry<br />

(Taphrina, Pro<strong>to</strong>myces). In Archiascomycetes,<br />

asci may (Taphrina, Pro<strong>to</strong>myces) or may not<br />

(Pneumocystis, Schizosaccharomyces) forcibly discharge<br />

their spores, whereas asci of Hemiascomycetes<br />

generally have evanescent walls, i.e.<br />

they release their ascospores passively.<br />

In the absence of asci and ascospores, the<br />

microscopic identification of yeasts is difficult<br />

or impossible, and other methods have <strong>to</strong><br />

be employed, e.g. physiological tests based on<br />

the ability of test species <strong>to</strong> grow on any of<br />

a standard set of carbon or nitrogen sources


262 HEMIASCOMYCETES<br />

(Yarrow, 1998; Barnett et al., 2000). The analysis of<br />

DNA sequences (e.g. 18S rDNA) is now performed<br />

routinely in many labora<strong>to</strong>ries, and a comparison<br />

with the extensive databases of appropriate<br />

sequences should afford identification at least <strong>to</strong><br />

genus level. In this way, hemiascomycete yeasts<br />

can be distinguished from Archiascomycetes<br />

and also from basidiomycete yeasts. Such a distinction<br />

should be unequivocal since it utilizes<br />

the very same characters by which the classes<br />

Hemi- and Archiascomycetes were established.<br />

10.1.1 Occurrence and isolation of<br />

Hemiascomycetes<br />

Hemiascomycete yeasts are prominent as epiphytic<br />

saprotrophic colonizers of plant organs,<br />

especially where sugars are present, e.g. in the<br />

nectar of flowers, on fruits, and on wounded<br />

or exposed surfaces of plants. Between 10 5 and<br />

10 7 yeast cells g 1 plant material (fresh weight)<br />

may be present (Phaff & Starmer, 1987). Yeasts<br />

also occur in the soil, although only a few exclusively<br />

soil-borne species have been described.<br />

Most yeasts are probably introduced in<strong>to</strong> the soil<br />

with the plant material with which they were<br />

originally associated (Phaff & Starmer, 1987).<br />

Yeasts also occur in freshwater and marine<br />

situations. Some species are associated with<br />

insects and other animals, including the guts of<br />

vertebrates which have a thriving yeast mycota.<br />

Yeasts may grow on skin surfaces and one<br />

species Candida albicans can, under certain<br />

circumstances, turn in<strong>to</strong> a mild or severe<br />

pathogen of humans, especially of immunocompromised<br />

patients (see p. 276). Hemiascomycetes<br />

are of little importance as plant pathogens with<br />

the exception of Eremothecium spp. which cause<br />

lesions on citrus fruits, cot<strong>to</strong>n and other<br />

crop plants, and are spread by sucking insects<br />

(see p. 284).<br />

Many species of Hemiascomycetes can grow<br />

under conditions of reduced water availability<br />

corresponding <strong>to</strong> about 50% glucose or a nearsaturated<br />

NaCl solution. Consequently, they can<br />

colonize most types of preserved foods, whereby<br />

the type of preservative determines the species<br />

composition (Pitt & Hocking, 1985; Fleet, 1990).<br />

Fortunately, food spoilage by yeasts does not<br />

normally result in the production of <strong>to</strong>xins,<br />

in contrast <strong>to</strong> bacteria or certain filamen<strong>to</strong>us<br />

fungi. However, the economic losses of food<br />

spoilage due <strong>to</strong> yeasts are still considerable.<br />

Hemiascomycete yeasts are easily isolated<br />

on<strong>to</strong> most standard agar media augmented<br />

with a suitable antibiotic <strong>to</strong> suppress bacteria,<br />

e.g. a mixture of penicillin G and strep<strong>to</strong>mycin<br />

sulphate (100 200 mg l 1 each), added <strong>to</strong> the<br />

cooling agar after au<strong>to</strong>claving. Plant or soil samples<br />

can be plated either directly, or the yeasts<br />

can be suspended by shaking the sample in<br />

sterile distilled water containing a detergent<br />

such as 0.01% (v/v) Tri<strong>to</strong>n X-100 or Tween 80.<br />

The undiluted sample or a dilution series in<br />

water can be plated out, and the density of<br />

colony-forming units (CFU) g 1 soil or leaves<br />

can be calculated. Yeasts are just large enough<br />

<strong>to</strong> be resolved as individual cells when a Petri<br />

dish is inverted and viewed with a 10 objective,<br />

whereas bacterial cells are not resolved at that<br />

magnification.<br />

10.1.2 The importance of<br />

Hemiascomycetes<br />

A very small number of species is of immense<br />

importance <strong>to</strong> biotechnology, and an adequate<br />

discussion is beyond the scope of this book.<br />

Below is a mention of the most important<br />

aspects; some further applications and the<br />

yeast species involved have been summarized<br />

by J. F. T. Spencer et al. (2002).<br />

1. Alcoholic fermentation mainly by<br />

Saccharomyces cerevisiae. This is the oldest and yet<br />

still the most important area of biotechnology,<br />

with about 10 11 l of beer and 3 10 10 l of wine<br />

produced worldwide each year (Oliver, 1991;<br />

Kurtzman & Sugiyama, 2001). The discovery of<br />

alcoholic fermentations has been made several<br />

times independently in the his<strong>to</strong>ry of mankind.<br />

Details of fermentation processes are given in<br />

on pp. 274 276. Industrial alcohol (ethanol) is<br />

often obtained from fermentations of corn<br />

starch hydrolysate by S. cerevisiae, but there is<br />

an ongoing interest in using other yeasts<br />

(Pachysolen tamophilus, Pichia stipitis) for the


SACCHAROMYCES (SACCHAROMYCETACEAE)<br />

263<br />

production of ethanol from pen<strong>to</strong>se sugars<br />

in wastes of industrial processes (Jeffries &<br />

Kurtzman, 1994).<br />

2. Bread-making. About 1.5 10 6 <strong>to</strong>ns of<br />

fresh cells of S. cerevisiae are produced worldwide<br />

per annum for use in the production of bread<br />

dough (see p. 274).<br />

3. Single-cell protein (SCP). This term describes<br />

the conversion of low-cost substrates in<strong>to</strong> proteinrich<br />

biomass of unicellular organisms. Yeasts<br />

have a high nutritional value <strong>to</strong> animals and<br />

man because they are rich in vitamins and<br />

protein, and because they do not generally<br />

produce myco<strong>to</strong>xins. Since they also have very<br />

simple growth requirements, yeasts can be used<br />

<strong>to</strong> convert low-cost substrates such as wastes from<br />

industrial processes in<strong>to</strong> high-value products for<br />

human or animal consumption. While the use of<br />

mineral oil as a substrate was a somewhat<br />

predictable failure, other substrates such as<br />

whey wastes from cheese production, molasses<br />

from sugar cane or pen<strong>to</strong>se-containing wastes<br />

from paper production are promising (Tuse, 1984;<br />

Scrimshaw & Murray, 1995; Paul et al., 2002).<br />

Currently, about 800 000 <strong>to</strong>ns of fodder yeasts are<br />

produced per annum (Kurtzman & Sugiyama,<br />

2001), but an extended application of single-cell<br />

protein technology is hampered by the low<br />

current cost of alternative protein sources such<br />

as soy meal or fish meal (Harrison, 1993;<br />

Scrimshaw & Murray, 1995).<br />

4. Vitamin production. Riboflavin (vitamin<br />

B2) is produced industrially by Eremothecium<br />

spp. (see p. 284).<br />

5. Production of recombinant proteins,<br />

e.g. enzymes, or clinically relevant molecules<br />

such as antigens, insulin and epidermal growth<br />

fac<strong>to</strong>r. Expression systems for heterologous<br />

proteins, i.e. proteins of interest whose gene<br />

has been linked <strong>to</strong> the promoter sequence of<br />

the producing organism, include S. cerevisiae<br />

and Pichia pas<strong>to</strong>ris. The latter holds advantages<br />

because proteins of interest are secreted more<br />

efficiently. Further, the glycosylation (sugar)<br />

chains which are added <strong>to</strong> the polypeptide<br />

during and after its translation in the rough<br />

endoplasmic reticulum are more similar between<br />

Pichia and mammals than either is <strong>to</strong> S. cerevisiae.<br />

Therefore, Pichia proteins cause fewer immunological<br />

problems in clinical use (see p. 281).<br />

6. Biological control. Because yeasts do not<br />

produce myco<strong>to</strong>xins and because of their ability<br />

<strong>to</strong> colonize the skin of fruits, they are being<br />

developed as biological control agents against<br />

postharvest losses in fruit crops. Pichia guilliermondii<br />

sprayed on<strong>to</strong> fruits selectively reduces<br />

development of moulds caused by Penicillium<br />

spp. (Chalutz & Wilson, 1990; McLaughlin et al.,<br />

1990).<br />

10.2 Saccharomyces<br />

(Saccharomycetaceae)<br />

About 10 16 species of Saccharomyces are<br />

currently recognized (Vaughan-Martini &<br />

Martini, 1998b; Barnett et al., 2000; Kirk et al.,<br />

2001). We will focus on S. cerevisiae, which in many<br />

ways is the most important fungus yet discovered.<br />

About 25 strains of S. cerevisiae exist, and these<br />

have different physiological properties which are<br />

relevant <strong>to</strong> their biotechnological applications.<br />

Many were formerly regarded as different species<br />

(Vaughan-Martini & Martini, 1998b; Rainieri<br />

et al., 2003). Saccharomyces cerevisiae is the brewer’s<br />

and baker’s yeast (see below), although some of<br />

the best brewing yeasts in current use belong<br />

<strong>to</strong> S. pas<strong>to</strong>rianus (¼ S. carlsbergensis). In nature,<br />

S. cerevisiae is found on ripe fruits, like many<br />

other yeasts. Grape and fruit wines are still often<br />

made by relying on spontaneous fermentations<br />

by yeasts which happen <strong>to</strong> be growing on the<br />

skin of the fruits used.<br />

The relatively small size of yeast cells<br />

(about 6 8 5 6 mm) has limited their investigation<br />

by light microscopy, but great progress<br />

has been made recently by the use of fluorescent<br />

dyes. Further, by fusing the green fluorescent<br />

protein (GFP) gene <strong>to</strong> the promoters of diverse<br />

yeast proteins, it has become possible <strong>to</strong> locate<br />

the site of a defined gene product within the<br />

yeast cell (Kohlwein, 2000). The availability<br />

of freeze-substitution fixation for transmission<br />

electron microscopy has led <strong>to</strong> the production<br />

of highly resolved ‘natural’ images of


264 HEMIASCOMYCETES<br />

organelles and cellular processes (Baba & Osumi,<br />

1987), although many issues of yeast cy<strong>to</strong>logy<br />

have remained controversial. A vegetative yeast<br />

cell (Fig. 10.1) is densely packed with organelles,<br />

including one nucleus, a large central<br />

vacuole, and mi<strong>to</strong>chondria. Mi<strong>to</strong>chondria are<br />

very dynamic in shape, showing a pronounced<br />

tendency <strong>to</strong> fuse in<strong>to</strong> one or a few large reticulate<br />

organelles when the energy demand is<br />

high, or <strong>to</strong> fragment in<strong>to</strong> numerous small<br />

promi<strong>to</strong>chondria during anaerobic fermentation<br />

or metabolic inactivity (Jensen et al., 2000;<br />

Okamo<strong>to</strong> & Shaw, 2005).<br />

An immense amount of literature exists on<br />

S. cerevisiae, covering aspects of genetics and<br />

molecular biology (Pringle et al., 1997), physiology<br />

(see Jennings, 1995), cy<strong>to</strong>logy (Baba & Osumi,<br />

1987) and biotechnology (Spencer & Spencer,<br />

1990; Walker, 1998). We can only broach a few<br />

selected <strong>to</strong>pics here <strong>to</strong> give an impression of<br />

Fig10.1 Saccharomyces cerevisiae. Sketch of a budding yeast cell as seen by transmission electron microscopy using material fixed by<br />

freeze-substitution.The presence of a morphologically recognizable Golgi stack is unusual among Eumycota. Modified and redrawn<br />

from Baba and Osumi (1987).


SACCHAROMYCES (SACCHAROMYCETACEAE)<br />

265<br />

the enormous importance of S. cerevisiae for<br />

fundamental cell biology. This fungus was the<br />

first eukaryote (in 1996) <strong>to</strong> have its complete<br />

genome sequenced, and this <strong>to</strong>gether with the<br />

ease of genetic manipulation has further<br />

enhanced the status of S. cerevisiae as a workhorse,<br />

if not ‘model organism’, for eukaryote<br />

research.<br />

10.2.1 The life cycle of S. cerevisiae<br />

Vegetative cells of S. cerevisiae are generally<br />

diploid in nature, although tetraploid or aneuploid<br />

strains also occur. Strains may be homoor<br />

heterothallic. The chromosome number is 16<br />

(Cherry et al., 1997). The life cycle of S. cerevisiae<br />

is presented in Figs. 10.2 and 10.3. The haploid<br />

ascospores often fuse within the ascus where<br />

they were formed (Fig. 10.3d), or shortly after<br />

release. However, if individual ascospores<br />

become isolated, they can germinate and reproduce<br />

as haploid cells by budding. Where two<br />

haploid cells of opposite mating type are in close<br />

contact, they secrete peptide hormones and produce<br />

plasma membrane-bound recep<strong>to</strong>rs which<br />

recognize the hormone of opposite mating type.<br />

The binding of a hormone molecule <strong>to</strong> the<br />

matching recep<strong>to</strong>r sets a signalling chain in<br />

motion (reviewed by Bardwell, 2004) which<br />

arrests the mi<strong>to</strong>tic cell cycle at G1, stimulates<br />

transcription of mating-specific genes and<br />

causes polarized growth of the two cells <strong>to</strong>wards<br />

each other (Leberer et al., 1997). Mating initially<br />

involves an increased ability of the surfaces of<br />

two cells <strong>to</strong> adhere <strong>to</strong> each other. This so-called<br />

sexual agglutination is mediated by glycoproteins.<br />

It seems that these are components of<br />

fimbriae, i.e. long filaments radiating outwards<br />

from the cell wall (see Fig. 23.15). Agglutination<br />

is followed by co-ordinated digestion of the<br />

walls separating the two cells. Plasmogamy and<br />

karyogamy follow swiftly (Gammie et al., 1998).<br />

The resulting diploid cell (Fig. 10.3e) can carry on<br />

reproducing asexually by budding. In contrast<br />

<strong>to</strong> Schizosaccharomyces pombe, there are therefore<br />

two mi<strong>to</strong>tic cycles in the life cycle of S. cerevisiae.<br />

Under optimum conditions, the culture doubling<br />

time by mi<strong>to</strong>sis is about 100 min.<br />

Diploid strains of S. cerevisiae can be induced<br />

<strong>to</strong> form ascospores by suitable treatment, and<br />

this yeast is therefore termed an ascosporogenous<br />

yeast, in contrast <strong>to</strong> asporogenous yeasts<br />

in which ascospores have not been observed.<br />

Meiosis can be induced by growing the yeast on<br />

a nutrient-rich presporulation medium containing<br />

an assimilable sugar, a suitable nitrogen<br />

source for good growth (nitrate is not utilized),<br />

Fig10.2 The life cycle of S. cerevisiae. Both haploid and diploid cells can reproduce by budding.Open and closed circles represent<br />

haploid nuclei of opposite mating type; diploid nuclei are larger and half-filled. Key events in the life cycle are plasmogamy (P),<br />

karyogamy (K) and meiosis (M).


266 HEMIASCOMYCETES<br />

Fig10.3 Saccharomyces cerevisiae. (a) Vegetative yeast cells<br />

reproducing by budding. (b) Yeast asci containing mostly four<br />

spores, sometimes with only three spores in focus. (c) Ascus<br />

showing a budding ascospore (arrow). (d) Ascus in which two<br />

spores have fused <strong>to</strong>gether and are budding. (e) Two ascospores<br />

fusing (<strong>to</strong>p left), and two fused ascospores forming a<br />

diploid bud (right).<br />

and vitamins of the B group. Such a medium<br />

results in well-grown cells which will sporulate<br />

on transfer <strong>to</strong> a sporulation medium. Sporulation<br />

occurs best on media in which budding<br />

is inhibited. Low concentrations of an assimilable<br />

carbon source are necessary <strong>to</strong> provide<br />

energy for the sporulation process. Acetate agar<br />

(1 g glucose, 8:2 g Na acetate3H 2 O and 2.5 g yeast<br />

extract l 1 ) is a good sporulation medium<br />

(Yarrow, 1998).<br />

Diploid yeast cells convert directly in<strong>to</strong><br />

asci within 12 24 h of incubation in a suitable<br />

sporulation medium. The frequency of ascus<br />

formation in most isolates is quite low, usually<br />

less than 10% (Vaughan-Martini & Martini,<br />

1998b). The cy<strong>to</strong>plasm differentiates in<strong>to</strong> four<br />

thick-walled spherical spores, although the<br />

number of spores may be fewer (see Fig. 10.3d).<br />

The nuclear divisions which precede ascus<br />

formation are meiotic. As is also the case in<br />

mi<strong>to</strong>sis, the nuclear envelope remains intact<br />

during meiosis, taking up a lobed shape as the<br />

nuclear spindles draw the chromosomes in<strong>to</strong> two<br />

and then four corners of the envelope (Fig. 10.4).<br />

The mechanism of ascospore formation has been<br />

extensively reviewed by Neiman (2005). It is very<br />

similar <strong>to</strong> that of higher ascomycetes, the main<br />

difference being that there is no common vesicle<br />

enclosing all nuclei prior <strong>to</strong> ascospore delimitation.<br />

Instead, a cup-shaped double-membrane,<br />

the prospore membrane, associates with each of<br />

the four spindle-pole bodies, and this gradually<br />

encapsulates its nuclear lobe until the four<br />

nuclei separate (Figs. 10.4g,h). As in other ascomycetes,<br />

the ascospore wall is then laid down<br />

in the lumen between the two membranes<br />

surrounding the developing ascospore.


SACCHAROMYCES (SACCHAROMYCETACEAE)<br />

267<br />

Fig10.4 Saccharomyces cerevisiae.<br />

Diagrammatic summary of the processes of<br />

meiosis and ascospore delimitation (from<br />

Beckett et al.,1974). (a d) The spindle pole<br />

body replicates and the two new spindle pole<br />

bodies move <strong>to</strong> opposite poles of the nucleus.<br />

The nuclear membrane remains intact. (e,f)<br />

Further replication of the spindle pole bodies<br />

and rearrangement.The nuclear envelope is<br />

still intact. New membranes, the ascosporedelimiting<br />

membranes, form outside the spindle<br />

pole bodies. (g,h) Envelopment of the<br />

lobes of the dividing nucleus by the ascospore-delimiting<br />

membranes results in the<br />

formation of four haploid uninucleate<br />

ascospores.<br />

10.2.2 Mating in S. cerevisiae<br />

Many strains of S. cerevisiae are heterothallic,<br />

and the ascospores are of two mating types.<br />

Mating type specificity is controlled by a single<br />

genetic locus which exists in two allelic states,<br />

a and a, and segregation at the meiosis preceding<br />

ascospore formation results in two a and two a<br />

ascospores. Fusion normally occurs only between<br />

cells of opposite mating type, and this has<br />

been termed legitimate copulation. Such fusions<br />

result in diploid cells which can, under appropriate<br />

conditions, form asci with viable<br />

ascospores.<br />

The mating type (MAT) alleles are rather small<br />

and structurally similar <strong>to</strong> each other, but differ<br />

in their central region which comprises about<br />

650 base pairs (bp) in MATa and 750 bp in MATa<br />

(Fig. 10.5). Because of this difference, mating type<br />

alleles are often called idiomorphs. In haploid<br />

a-cells, the MATa locus expresses two genes, both<br />

of which encode regula<strong>to</strong>ry proteins. The a1 gene<br />

product interacts with a constitutively expressed<br />

protein not encoded by the MAT locus, Mcm1p,<br />

<strong>to</strong> activate several a-specific genes outside<br />

the MAT locus, notably those encoding the<br />

a-pheromone which is secreted by a-cells, and


268 HEMIASCOMYCETES<br />

Fig10.5 The structure of the mating type idiomorphs a (<strong>to</strong>p) and a (bot<strong>to</strong>m) of Saccharomyces cerevisiae.The two alleles differ only<br />

in their central (Y) regions which contain parts of two genes a1anda2ora1anda2.The entire lengths of these genes and their<br />

directions of transcription are indicated by arrows.The function of a2 is unknown. In diploid cells, the lack of expression of a1andthe<br />

formation of a dimeric a2/a1protein suppresses the expression of mating type-specific proteins including hormones and their<br />

recep<strong>to</strong>rs. nt ¼ nucleotides.Redrawn from Haber (1998) Annual Reviews of Genetics 32,withpermission.ß1998 Annual Reviews,<br />

www.annualreviews.org.<br />

the plasma membrane recep<strong>to</strong>r Ste2p, which<br />

can bind a-pheromone from the environment.<br />

In diploid cells, expression of a1 and thus of<br />

a-specific proteins is repressed. The a2 gene<br />

encodes a repressor protein which interacts<br />

with several other regula<strong>to</strong>ry proteins,<br />

including Mcm1p, <strong>to</strong> repress the expression of<br />

a-specific genes, including those encoding the<br />

a-pheromone and the a-fac<strong>to</strong>r recep<strong>to</strong>r Ste3p. In<br />

the absence of the a1 and a2 gene products,<br />

haploid cells have an a-phenotype with respect <strong>to</strong><br />

mating behaviour because a-genes are constitutively<br />

expressed. The function of a2 is unknown,<br />

and the a1 gene product is active only in diploid<br />

cells, combining with the a2 protein <strong>to</strong> repress<br />

haploid-specific genes including those encoding<br />

the two pheromones and their recep<strong>to</strong>rs.<br />

Another gene repressed by the a2/a1 dimer<br />

is RME1, which encodes a repressor of meiosis.<br />

Meiosis can therefore only take place if a diploid<br />

a/a nucleus exists in which Rme1p is repressed<br />

by the a2/a1 dimer, and if nutrient conditions<br />

are limiting. The signal for nutrient limitation<br />

is sensed and transmitted by a cyclic AMPdependent<br />

signalling chain (Klein et al., 1994).<br />

Therefore, the most important difference<br />

between a/a diploids and homozygous diploids<br />

or haploids is that only the a/a diploids can initiate<br />

meiosis under nutrient-limiting conditions,<br />

leading <strong>to</strong> the production of ascospores which<br />

are more resistant <strong>to</strong> adverse conditions than<br />

vegetative cells. It is likely that this enhanced<br />

survival of ascospores is the reason why haploid<br />

or homozygous populations of S. cerevisiae and<br />

some other ascomycetes (e.g. Schizosaccharomyces<br />

pombe) possess the intriguing ability <strong>to</strong> switch<br />

their mating type, thereby acquiring the ability<br />

<strong>to</strong> undergo meiosis. An excellent account of the<br />

experiments and ideas leading <strong>to</strong> the unravelling<br />

of the mating fac<strong>to</strong>r switch in S. cerevisiae has<br />

been given by Haber (1998), and we borrow<br />

heavily from it in the following summary.<br />

Strathern and Herskowitz (1979) observed<br />

that the ability <strong>to</strong> switch their mating type is<br />

acquired only by cells which have previously<br />

divided at least once. The pattern established by<br />

a single germinating a-type haploid ascospore is<br />

shown in Fig. 10.6: the first daughter cell has the<br />

mating type a, but then the mother cell switches<br />

its mating type prior <strong>to</strong> its second division,<br />

so that two a-cells result. Meanwhile, the first<br />

daughter (a-type) cell undergoes its first division<br />

so that a cluster of two a- and two a-type cells<br />

is produced. Conjugation can occur, and the<br />

two zygotes can carry on dividing as diploid<br />

yeast cells with the additional option <strong>to</strong><br />

undergo meiosis and produce asci if required.<br />

The mating type switch is brought about by the


SACCHAROMYCES (SACCHAROMYCETACEAE)<br />

269<br />

Fig10.6 Mating type switch in Saccharomyces<br />

cerevisiae. A germinating a-type ascospore<br />

denoted by its white nucleus produces a bud<br />

which has the same mating type. Ash1mRNA is<br />

selectively translocatedin<strong>to</strong> the bud (a), and its<br />

protein Ash1p is expressed in the daughter<br />

nucleus (b), preventing it from switching its<br />

mating type. However, the mother cell is now<br />

competent <strong>to</strong> switch its mating type <strong>to</strong> a<br />

because it has divided once before (c); when it<br />

undergoes its second division (d,e), it produces<br />

an a-type daughter cell.The<br />

first-formed daughter cell cannot switch its<br />

mating type because it has no previous his<strong>to</strong>ry<br />

of cell division.Consequently it produces a<br />

daughter of a-type. Mating (f) occurs between<br />

one a-andonea-type cell apiece, and conjugation<br />

results in two diploid cells which can<br />

reproduce by further budding (g) or by meiosis<br />

and ascus formation. Based on Haber (1998).<br />

HO endonuclease, the gene of which is only<br />

expressed in mother cells which have divided<br />

at least once (Nasmyth, 1983). Expression of the<br />

HO endonuclease gene is controlled by a series of<br />

repressor proteins, and one of them Ash1p is<br />

confined <strong>to</strong> the daughter cell upon division. This<br />

is due <strong>to</strong> the selective transport of its mRNA<br />

molecule in<strong>to</strong> the bud prior <strong>to</strong> cy<strong>to</strong>kinesis (Long<br />

et al., 1997).<br />

Hicks et al. (1977) proposed the cassette model<br />

<strong>to</strong> account for the mating type switch, and this<br />

has been confirmed by subsequent experimentation.<br />

In addition <strong>to</strong> the mating type locus which<br />

is active in a given haploid cell, each haploid<br />

genome possesses two further complete copies,<br />

one <strong>to</strong> the left of the active locus which usually<br />

contains the a allele (i.e. HMLa) and the other <strong>to</strong><br />

the right, encoding a (i.e. HMRa). These genes are<br />

silenced, i.e. they are not transcribed because<br />

their DNA is coated by his<strong>to</strong>nes and other proteins<br />

encoded and regulated by numerous other<br />

genes. Silencing is determined by the location<br />

of these gene copies in the proximity of silencing<br />

sequences (Haber, 1998). The mating type switch<br />

is perfomed when the HO endonuclease cleaves<br />

the currently active locus at the Y/Z boundary<br />

(see Fig. 10.5), followed by the digestion of one<br />

of the two DNA strands of the Z region by an<br />

exonuclease. A new sequence is then copied<br />

in<strong>to</strong> that gap from either of the two silent genes,<br />

using the one-stranded Z region as a template.<br />

The integrity of the silent gene which acts as


270 HEMIASCOMYCETES<br />

template is unaffected during that process<br />

(for details, see Haber, 1998). Hence, yeast cells<br />

can repeatedly switch their mating type. There is<br />

an element of selectivity in the mating type<br />

switch because, for example, a-cells choose the<br />

silent a-locus 85 90% of the time (Weiler &<br />

Broach, 1992). In the case of a-cells, preference<br />

for the switch <strong>to</strong> the a mating type is brought<br />

about by the a2 protein (for details, see Haber,<br />

1998).<br />

10.2.3 The cell wall of S. cerevisiae<br />

The wall of S. cerevisiae represents a considerable<br />

biochemical investment, making up 15 30% of<br />

the dry weight of vegetative cells. Up <strong>to</strong> three<br />

wall layers can be distinguished by electron<br />

microscopy. They differ in their chemical composition,<br />

and the relatively simple architecture of<br />

the wall of S. cerevisiae is considered a model for<br />

other fungi (Molina et al., 2000; de Nobel et al.,<br />

2001). The middle layer is electron-translucent<br />

and consists of the main structural scaffold<br />

of branched b-(1,3)-glucan molecules which bind<br />

b-(1,6)-glucans and chitin. The latter, however, is<br />

present only in low quantities (1 2% of the <strong>to</strong>tal<br />

wall material) and it is unevenly distributed,<br />

being concentrated in a ring around the region<br />

where the bud emerges. The outer wall layer of<br />

S. cerevisiae is electron-dense because it consists<br />

mainly of proteins. These determine the<br />

cell surface properties, including the porosity<br />

(pore size) of the cell wall (Zlotnik et al., 1984)<br />

and adhesiveness <strong>to</strong> other cells (flocculation; see<br />

p. 274). The outer wall proteins may be highly<br />

glycosylated in S. cerevisiae by the addition of<br />

large mannose chains. In pathogenic yeasts such<br />

as Candida albicans, this outer layer is also<br />

important because of its involvement in attachment<br />

of the fungus <strong>to</strong> its host, and because it<br />

conveys antigenic properties. There are two main<br />

groups of outer cell wall proteins (CWPs) in<br />

S. cerevisiae. The members of one are modified<br />

by a glycosylphosphatidylinosi<strong>to</strong>l (GPI) chain<br />

which is indirectly linked <strong>to</strong> the b-(1,3)-glucans<br />

of the central wall layer via the b-(1,6)-glucans.<br />

These proteins are called GPI CWPs. The second<br />

type of outer cell wall protein is called Pir CWP<br />

(Pir ¼ protein with internal repeats) and is<br />

linked directly <strong>to</strong> the b-(1,3)-glucan component<br />

(Kapteyn et al., 1999; de Nobel et al., 2001). Both<br />

GPI CWPs and Pir CWPs are structural<br />

proteins. The innermost layer (periplasmic<br />

space) is also electron-dense and consists of<br />

proteins, but these are mostly enzymes which<br />

are <strong>to</strong>o large <strong>to</strong> pass through the central layer.<br />

They are therefore restrained by the glucan layer<br />

(de Nobel & Barnett, 1991).<br />

The polarity of wall synthesis in S. cerevisiae<br />

is controlled by the localization of the plasma<br />

membrane-bound enzymes (glucan synthetases,<br />

chitin synthetases) which produce the elements<br />

of the middle layer, and by the secretion of the<br />

structural outer wall proteins as well as enzymes<br />

which cross-link the various elements of the<br />

cell wall. Cell wall synthesis is thus regulated<br />

spatially by the polarity of the yeast cell, and<br />

temporally by the cell cycle; the transcription<br />

of many genes involved in cell wall synthesis<br />

is cell cycle-dependent (Molina et al., 2000;<br />

Rodríguez-Peña et al., 2000).<br />

10.2.4 Morphogenesis and the cell cycle<br />

of S. cerevisiae<br />

One oddity about S. cerevisiae is that it does<br />

not require microtubules for the maintenance<br />

of its cellular polarity, as shown by mutant<br />

and inhibi<strong>to</strong>r studies. In other eukaryotes,<br />

including filamen<strong>to</strong>us fungi and also the fission<br />

yeast Schizosaccharomyces pombe, microtubules<br />

are employed for long-distance transport processes.<br />

It is possible that they are dispensible in<br />

S. cerevisiae simply because of the small distance<br />

between the mother cell and the growing<br />

bud. Of course, microtubules are required in<br />

S. cerevisiae as in all other eukaryotes for nuclear<br />

division. Actin, in contrast, is crucial for cell<br />

polarity and cell viability in S. cerevisiae (Pruyne<br />

& Bretscher, 2000b; Pruyne et al., 2004).<br />

The cell cycle of S. cerevisiae is similar <strong>to</strong> that<br />

of Schizosaccharomyces pombe (see p. 256) and other<br />

eukaryotes in its regula<strong>to</strong>ry mechanisms (Lewin,<br />

2000; Alberts et al., 2002), except that the G2<br />

phase is lacking, and that cell division (cy<strong>to</strong>kinesis)<br />

is initiated early in the cycle, the bud being<br />

already present during the S phase. A summary<br />

of the budding process is given in Fig. 10.7.


SACCHAROMYCES (SACCHAROMYCETACEAE)<br />

271<br />

Fig10.7 The mi<strong>to</strong>tic cell cycle of Saccharomyces cerevisiae.<br />

Actin patches and cables are drawn in white; the septin ring is<br />

black. Secre<strong>to</strong>ry vesicles are not drawn, but their distribution<br />

follows essentially that of actin patches. (a) Initiation of the<br />

bud site occurs during the S phase by the assembly of actin<br />

patches around a septin ring. (b) Bud extension during late S<br />

phase.The cap co-ordinates apical bud growth by establishing<br />

actin cables which mediate the transport of secre<strong>to</strong>ry vesicles<br />

in<strong>to</strong> the bud, and their fusion in the region of the cap. (c) A<br />

later stage of bud growth; the cap components become<br />

distributed more evenly over the bud membrane surface, and<br />

growth is non-polarized (isodiametric). Meanwhile, the dividing<br />

nucleus is drawn, via cy<strong>to</strong>plasmic microtubules (not<br />

shown), <strong>to</strong> an actin/myosin ring superimposed on<strong>to</strong> the septin<br />

ring.The nucleus divides so that each cell receives one daughter<br />

nucleus. (d) Re-establishment of polarized growth by the<br />

formation of two septin rings, one on either side of the bud<br />

site.Growth leads <strong>to</strong> closure of the pore between the mother<br />

and daughter cell. Redrawn from Sheu and Snyder (2001), with<br />

kind permission of Springer Science and Business Media.<br />

Useful and comprehensive reviews are those by<br />

Pruyne and Bretscher (2000a,b) and Sheu and<br />

Snyder (2001). The bud site is determined by the<br />

assembly of a protein cap at the inner surface<br />

of the plasma membrane. This cap contains an<br />

essential regula<strong>to</strong>ry protein, Cdc42p, which is<br />

controlled directly by the cell cycle and in turn<br />

determines the sequestration of numerous other<br />

scaffold and regula<strong>to</strong>ry proteins by the cap.<br />

Budding differs between haploid and diploid<br />

cells, the former initiating new buds adjacent<br />

<strong>to</strong> the previous one, and the latter budding in<br />

a bipolar fashion (Pruyne et al., 2004).<br />

One important group of proteins are the<br />

septins, which form a ring around the bud site<br />

(Fig. 10.7a). Actin filaments are initiated from<br />

the centre of the cap. As the bud extends in a<br />

polarized fashion, the septin ring remains at the<br />

site of bud emergence whereas the cap which<br />

governs bud extension migrates with the bud,<br />

staying at the apex and controlling bud extension<br />

(Fig. 10.7b). Later, the cap components and<br />

actin filament attachment points become distributed<br />

diffusely over the bud surface. This<br />

leads <strong>to</strong> the fusion of secre<strong>to</strong>ry vesicles over<br />

the entire bud surface, and <strong>to</strong> a change in the<br />

growth pattern from polarized <strong>to</strong> isodiametric<br />

(Fig. 10.7c).<br />

The septin ring binds numerous proteins,<br />

including important regula<strong>to</strong>ry ones (Versele &<br />

Thorner, 2005). Actin and myosin are attracted<br />

early during bud emergence, and a contractile<br />

actin ring is superimposed on the septin ring.<br />

Later, during nuclear division, cy<strong>to</strong>plasmic<br />

microtubules are also captured; these, in turn,<br />

are in contact with the microtubules of the<br />

intranuclear mi<strong>to</strong>tic spindle, and thus the dividing<br />

nucleus is drawn <strong>to</strong>wards the ring, with one<br />

daughter nucleus apiece ending up in each of<br />

the two cells (Fig. 10.7c; Kusch et al., 2002).<br />

Cy<strong>to</strong>kinesis is brought about as the actin ring<br />

contracts (Lippincott & Li, 1998). At this point,<br />

septins appear as a double ring, sandwiching<br />

the constricting actin ring. The septin double<br />

ring assembles two caps, and these serve as a<br />

nucleation centre for actin filaments which<br />

direct secre<strong>to</strong>ry vesicles <strong>to</strong> the bud site, closing<br />

the wall between mother and daughter cell<br />

(Fig. 10.7d). The regula<strong>to</strong>ry mechanisms are<br />

immensely complex but are beginning <strong>to</strong> be<br />

unravelled (Pruyne & Bretscher, 2000a,b;<br />

Pruyne et al., 2004; Versele & Thorner, 2005)<br />

and are likely <strong>to</strong> be of fundamental significance<br />

because the division of cells by a constricting<br />

actin ring is found also in other fungi and<br />

in animal systems, although apparently not in<br />

plants.


272 HEMIASCOMYCETES<br />

Following separation of the daughter cell,<br />

a circular, crater-like bud scar is left as a<br />

permanent mark on the surface of the mother<br />

cell (see Fig. 10.12). The maximum number of<br />

scars that could be accommodated on the surface<br />

of a yeast cell is about 100, suggesting that<br />

individual yeast cells are not capable of unlimited<br />

budding. Individual yeast cells age just like<br />

other organisms, although the timing of death<br />

is determined by complicated genetic fac<strong>to</strong>rs<br />

and the sum of metabolic energy expended<br />

throughout the life of the yeast cell, rather<br />

than the number of the bud scars per se<br />

(Jazwinski, 2002).<br />

Under certain environmental conditions<br />

(notably nutrient deficiency) diploid and, <strong>to</strong> a<br />

lesser extent, haploid cells of S. cerevisiae can<br />

change their growth pattern from budding,<br />

which produces heaps of cells only on the agar<br />

surface, <strong>to</strong> the formation of pseudohyphae which<br />

can grow in<strong>to</strong> the agar. Pseudohyphae may be<br />

of significance in the ecology of S. cerevisiae<br />

because they allow the organism <strong>to</strong> spread over<br />

and penetrate in<strong>to</strong> substrates, and <strong>to</strong> assimilate<br />

nutrients more readily (Gimeno et al.,<br />

1992). Formation of pseudohyphae requires an<br />

enhanced adhesion of the cells <strong>to</strong> each other and<br />

an enhanced polarity of daughter cell growth.<br />

Not surprisingly, the signalling events leading<br />

<strong>to</strong> pseudohyphal growth are rather complex<br />

(Palecek et al., 2002; Cecca<strong>to</strong>-An<strong>to</strong>nini &<br />

Sudbery, 2004).<br />

10.2.5 Membrane cycling in S. cerevisiae<br />

An enormous amount of work has been done<br />

<strong>to</strong> elucidate the secre<strong>to</strong>ry route in S. cerevisiae,<br />

and a sizeable collection of temperature-sensitive<br />

mutants with defects at different points of the<br />

secre<strong>to</strong>ry route has been assembled (Schekman,<br />

1992). Further, individual stepwise modifications<br />

<strong>to</strong> proteins travelling the secre<strong>to</strong>ry route can be<br />

identified, especially with respect <strong>to</strong> their<br />

glycosylation pattern and proteolytic cleavage<br />

of parts of the original polypeptide chain<br />

(Graham & Emr, 1991). The export of proteins<br />

starts with their synthesis in the rough endoplasmic<br />

reticulum and continues with their<br />

processing in a Golgi system. Along this route,<br />

the proteins are modified by the addition of<br />

glycosylation chains, and by the proteolytic<br />

cleavage of signal sequences. Transport has<br />

long been thought <strong>to</strong> occur by means of vesiclelike<br />

carriers, and the biochemical events leading<br />

<strong>to</strong> the budding of a vesicle from its source<br />

and its fusion with the destination membrane<br />

(e.g. ER ! Golgi) have been extensively characterized<br />

(Rothman & Orci, 1992). However, it is<br />

still unclear whether discrete vesicular carriers<br />

are an obligate transport system in vegetative<br />

yeast cells. An alternative is the dynamic maturation<br />

model in which sheets of ER become<br />

transformed in<strong>to</strong> Golgi compartments which<br />

gradually dilate and fragment in<strong>to</strong> secre<strong>to</strong>ry<br />

vesicles (Rambourg et al., 2001). Whatever their<br />

initial his<strong>to</strong>ry, secre<strong>to</strong>ry vesicles emerge from<br />

the Golgi system (Baba & Osumi, 1987) and<br />

migrate <strong>to</strong> the growing bud along actin cables<br />

(Finger & Novick, 1998).<br />

The cell membrane shows a high capacity<br />

for endocy<strong>to</strong>sis, i.e. the removal of excess membrane<br />

material and the uptake of specific molecules<br />

by membrane-bound recep<strong>to</strong>rs from the<br />

liquid medium of the environment. The occurrence<br />

of endocy<strong>to</strong>sis has been controversial<br />

in filamen<strong>to</strong>us fungi, but it has been obvious<br />

for some time that this must take place in<br />

S. cerevisiae as it is the route through which<br />

mating hormones are internalized and transported<br />

<strong>to</strong> the vacuole for degradation. While<br />

actin is certainly involved in endocy<strong>to</strong>sis, it is<br />

still unclear whether the actin patches long<br />

known <strong>to</strong> exist inside the plasma membrane of<br />

S. cerevisiae cells are the scaffold around which<br />

the inward-budding of the plasma membrane is<br />

moulded (Shaw et al., 2001). Endocy<strong>to</strong>sis occurs<br />

when pits are formed at the plasma membrane<br />

and bud inwards <strong>to</strong> form small vesicles (endocy<strong>to</strong>tic<br />

vesicles) which fuse <strong>to</strong> form a tubular<br />

early endosome. From there, material is transported<br />

via a late endosome <strong>to</strong> the vacuole in<br />

which it is degraded (Munn, 2000; Shaw et al.,<br />

2001). The protein ubiquitin plays a vital role as<br />

a tag for endocy<strong>to</strong>sis at the plasma membrane<br />

and for transport of endosomes <strong>to</strong> the vacuole<br />

(Horák, 2003). The purposes of endocy<strong>to</strong>sis could<br />

include the removal of excess membrane material,<br />

the removal of nutrient uptake systems


SACCHAROMYCES (SACCHAROMYCETACEAE)<br />

273<br />

no longer required, and the removal of mating<br />

type recep<strong>to</strong>rs, e.g. after the fusion of two<br />

haploid cells.<br />

10.2.6 The yeast vacuole<br />

The vacuole is the central destination of membrane<br />

trafficking in S. cerevisiae. It receives input<br />

directly from the secre<strong>to</strong>ry route in the form<br />

of most vacuolar proteins which are separated<br />

at the Golgi stage from those bound for secretion.<br />

Material also reaches the vacuole from<br />

the endocy<strong>to</strong>tic route (see above), and from<br />

the cy<strong>to</strong>plasm, especially during starvation.<br />

Cy<strong>to</strong>plasmic material may be engulfed directly<br />

by the <strong>to</strong>noplast (microau<strong>to</strong>phagy) or redundant<br />

material may first be surrounded by a double<br />

membrane <strong>to</strong> form an au<strong>to</strong>phagosome whose<br />

outer membrane then fuses with the <strong>to</strong>noplast<br />

(macroau<strong>to</strong>phagy). Details of these processes<br />

have been described by Klionsky (1997) and<br />

Thumm (2000). Degradation of protein is a<br />

major function of the vacuole in starvation<br />

situations, and about 40% of the <strong>to</strong>tal protein<br />

content of a yeast cell can be degraded<br />

within 24 h (Teichert et al., 1989). Not surprisingly,<br />

the vacuole contains a large set of powerful<br />

hydrolytic enzymes, especially proteases<br />

(Klionsky et al., 1990).<br />

When nitrogen is abundant, it is s<strong>to</strong>red in<br />

the vacuole as arginine at concentrations of<br />

up <strong>to</strong> 400 mM, and this can be re-released in<strong>to</strong><br />

the cy<strong>to</strong>plasm if nitrogen becomes limiting<br />

(Kitamo<strong>to</strong> et al., 1988). Likewise, phosphate can<br />

be s<strong>to</strong>red and released (Castro et al., 1999), as<br />

can many other ionic nutrients (Jennings, 1995).<br />

Toxic ions and metabolites may be s<strong>to</strong>red in<br />

the vacuole (e.g. Ramsay & Gadd, 1997). Vacuoles<br />

thus fulfil a crucial function in maintaining<br />

the homeostasis of the yeast cy<strong>to</strong>plasm against<br />

changing external conditions. In order <strong>to</strong> fulfil<br />

such functions, the vacuolar morphology<br />

can change dramatically, e.g. by fragmentation<br />

of one large central vacuole in<strong>to</strong> numerous small<br />

ones (Çakar et al., 2000).<br />

10.2.7 Killer yeasts and killer <strong>to</strong>xins<br />

Killer yeasts are strains which produce <strong>to</strong>xins<br />

capable of killing other strains belonging <strong>to</strong> the<br />

same or <strong>to</strong> closely related species. Toxin producers<br />

are resistant against their own <strong>to</strong>xin, but may<br />

be susceptible <strong>to</strong> <strong>to</strong>xins produced by other<br />

strains. Three important virus-encoded killer<br />

<strong>to</strong>xins (K1, K2, K28) are known <strong>to</strong> exist in<br />

S. cerevisiae; all three are polypeptides and are<br />

encoded by double-stranded RNA encapsulated<br />

in virus-like particles (VLPs). Another group of<br />

double-stranded viruses (the L-A viruses) belonging<br />

<strong>to</strong> the genus Totivirus is necessary for the<br />

replication of the killer <strong>to</strong>xin VLPs. The subject<br />

of killer yeasts has been reviewed by Magliani<br />

et al. (1997) and Marquina et al. (2002).<br />

The best-researched killer <strong>to</strong>xin is K1. It is<br />

encoded by a single open reading frame and<br />

is synthesized as a single polypeptide which is<br />

initially localized in the ER membrane. As the<br />

membrane-bound polypeptide travels the secre<strong>to</strong>ry<br />

route, it is modified by glycosylation and<br />

proteolytic cleavage, much like other secre<strong>to</strong>ry<br />

proteins. In the Golgi system, the polypeptide is<br />

cleaved in<strong>to</strong> two parts which are held <strong>to</strong>gether<br />

by disulphide bonds, and a third part, the glycosylated<br />

region, which is not part of the active<br />

<strong>to</strong>xin. The active <strong>to</strong>xin is secreted and diffuses<br />

in<strong>to</strong> the growth medium. The two parts of the<br />

active molecule fulfil two different functions;<br />

the b-chain binds the molecule <strong>to</strong> its recep<strong>to</strong>r<br />

site which is the b-(1,6)-glucan component of<br />

the cell wall. Following binding <strong>to</strong> the wall, the<br />

<strong>to</strong>xin is thought <strong>to</strong> be transferred <strong>to</strong> the<br />

plasma membrane where the a-chain forms<br />

a trans-membrane pore. Death of the target cell<br />

occurs because the trans-plasma membrane<br />

pro<strong>to</strong>n and ionic gradients are disrupted. The<br />

<strong>to</strong>xin can bind <strong>to</strong> the wall of the producing<br />

cell but not <strong>to</strong> its plasma membrane; presumably<br />

a membrane recep<strong>to</strong>r is altered, masked<br />

or destroyed. Self-immunity is conveyed by a<br />

precursor molecule of the mature <strong>to</strong>xin (Boone<br />

et al., 1986).<br />

The K1 <strong>to</strong>xin has been an important instrument<br />

in elucidating the processing of proteins<br />

along the secre<strong>to</strong>ry route, and the mechanism<br />

of cell wall synthesis in S. cerevisiae. Additionally,<br />

there are biotechnological implications. The<br />

possession of a killer <strong>to</strong>xin conveys a selective<br />

advantage upon a yeast strain, and killer yeasts<br />

are particularly common (25% of all isolates)


274 HEMIASCOMYCETES<br />

in habitats which contain abundant nutrients,<br />

such as the surface of ripe fruits (Starmer<br />

et al., 1987). Not surprisingly, contaminations<br />

by killer yeasts can be a problem in long-term<br />

fermentation processes, e.g. wine production<br />

(van Vuuren & Jacobs, 1992), and attempts have<br />

been made <strong>to</strong> incorporate the killer virus in<strong>to</strong><br />

yeast strains used for biotechnological purposes<br />

(Javadekar et al., 1995). Killer <strong>to</strong>xins<br />

cannot themselves be used against clinically<br />

relevant yeasts because the molecules are so<br />

large that they would elicit an immune response<br />

in the patient. However, it is possible <strong>to</strong> create<br />

antibodies which mimic the membrane-disrupting<br />

action of killer <strong>to</strong>xins (Polonelli et al., 1991).<br />

Whether these will become useful in medicine,<br />

e.g. against Candida infections, remains <strong>to</strong> be<br />

seen, but the existence of natural human<br />

antibodies with a killer <strong>to</strong>xin effect on Candida<br />

points <strong>to</strong> potential applications of this strategy<br />

(Magliani et al., 1997, 2005).<br />

10.2.8 Bread-making<br />

The principle behind the leavening of bread<br />

dough by baker’s yeast is the same as in brewing,<br />

i.e. the anaerobic metabolism of glucose and<br />

other reducing sugars via pyruvic acid in<strong>to</strong><br />

ethanol and CO 2 . The difference is that the<br />

released CO 2 is the important product in breadmaking<br />

because it is responsible for the texture<br />

of the bread. Ethanol may, however, contribute<br />

<strong>to</strong> the flavour of fresh bread. Originally, a portion<br />

of the risen dough medium was retained as<br />

a starter for the next baking session, or surplus<br />

yeast from brewing processes was used (Jenson,<br />

1998). Specific yeasts for baking were first<br />

produced in Vienna in 1846, and baker’s yeast<br />

is now produced commercially under aerobic<br />

conditions because the yield of biomass can be<br />

maximized (Caron, 1995). In the bread dough the<br />

yeast cells are subjected <strong>to</strong> anoxic or anaerobic<br />

conditions and must be able <strong>to</strong> release CO 2<br />

quickly. The carbon sources available <strong>to</strong> yeast<br />

cells in bread dough are hexoses, especially glucose,<br />

and the disaccharides mal<strong>to</strong>se and sucrose,<br />

all of which are present at fairly low concentrations.<br />

Starch is not utilized by S. cerevisiae but can<br />

be hydrolysed by amylases present in the flour,<br />

and the glucose thereby released may be<br />

available <strong>to</strong> the yeast (Oliver, 1991; Jakobsen et<br />

al., 2002). A very thorough account of the<br />

microbiology and processes of baking is that<br />

by Spicher and Brümmer (1995).<br />

10.2.9 Beer brewing<br />

Several good accounts of the process of beer<br />

brewing have been given (e.g. Oliver, 1991;<br />

Russell & Stewart, 1995; Hartmeier & Reiss,<br />

2002), and there are numerous popular books<br />

exploring the diversity of beers worldwide.<br />

In his masterful his<strong>to</strong>ry of beer, Hornsey (2003)<br />

has summarized evidence of the first known<br />

records and recipes of beer which date back<br />

6000 years or more and originate from Mesopotamia<br />

and ancient Egypt, where beer was more<br />

widely consumed than wine. The art of brewing<br />

may be almost as ancient as the cultivation<br />

of cereals, and indeed some his<strong>to</strong>rians believe<br />

that brewing was a major incentive for the<br />

development of agriculture around 6000 BC<br />

(Hornsey, 2003). Brewing has remained the<br />

most important area of biotechnology <strong>to</strong> this<br />

day. As<strong>to</strong>nishing quantities of beer are being<br />

consumed, with several sources agreeing on the<br />

Czech Republic as the <strong>to</strong>p beer-drinking nation<br />

at around 160 l per person per annum, followed<br />

by the Republic of Ireland (155 l) and Germany<br />

(128 l). These values appear frugal when put<br />

in<strong>to</strong> the his<strong>to</strong>rical context, e.g. of one gallon<br />

(3.8 l) as the daily personal allowance of ale for<br />

monks in medieval England (Hornsey, 2003).<br />

Two fundamentally different types of fermentation<br />

exist, and these differ in the strains of<br />

yeast used. In bot<strong>to</strong>m-fermenting beers, especially<br />

lager beers, the yeast settles as a sludge at<br />

the bot<strong>to</strong>m of the brewing vessel at the end<br />

of the fermentation, whereas it floats at the <strong>to</strong>p<br />

in <strong>to</strong>p-fermenting beers, especially the English<br />

ales, porters, s<strong>to</strong>uts, and the German Altbier. In<br />

Germany, a purity law was passed in 1516 which<br />

banned the use of ingredients other than water,<br />

yeast, malted barley and hops. Although now<br />

formally abolished by the European Union, most<br />

brewers still abide by it. In contrast, ale and<br />

lager brewers outside Germany often add other<br />

ingredients <strong>to</strong> their beer, e.g. cereals other than<br />

barley, other fermentable sugar sources such


SACCHAROMYCES (SACCHAROMYCETACEAE)<br />

275<br />

as syrups or enzymatic starch digests, fruits or<br />

interesting spices.<br />

Since S. cerevisiae cannot utilize starch, this<br />

has <strong>to</strong> be hydrolysed in<strong>to</strong> fermentable sugars<br />

first. The starch reserves in the endosperm of<br />

barley are hydrolysed naturally by endogenous<br />

amylases when the grains germinate. At a<br />

suitable time, the germination process is terminated<br />

by drying and heating (kilning), and the<br />

degree <strong>to</strong> which the malt is roasted determines<br />

the colour of the beer. For instance, heavily<br />

roasted malts are used for the dark milds, porter<br />

ales and s<strong>to</strong>uts. During mashing, the ground<br />

malt is heated in water <strong>to</strong> about 65°C; surviving<br />

endogenous enzymes or added enzymes continue<br />

with starch hydrolysis, and with the<br />

degradation of proteins <strong>to</strong> amino acids. Hops<br />

are added <strong>to</strong> the liquid (wort) which is then<br />

boiled, traditionally in a copper vessel, and after<br />

filtration and cooling the yeast is added. Top<br />

fermentation of ales takes a few days at 15 22°C<br />

whereas lager beers are fermented for up <strong>to</strong><br />

2 weeks at 8 15°C. In addition <strong>to</strong> ethanol, yeast<br />

metabolites which impart flavour <strong>to</strong> the beer are<br />

esters of higher alcohols (e.g. isobutanol), dike<strong>to</strong>nes,<br />

diacetyl, isobutyraldehyde and methylglyoxal.<br />

Sulphur-containing metabolites may also<br />

be important. There are several ways in which<br />

the flavour profile can be modified <strong>to</strong> give a<br />

desirable taste. For example, a yeast strain<br />

with the appropriate ester profile can be used,<br />

or new strains expressing the required enzymes<br />

can be engineered. A re-fermentation, often<br />

with Brettanomyces spp., may be performed in<br />

order <strong>to</strong> alter the flavour profile (Vanderhaegen<br />

et al., 2003).<br />

Towards the end of the fermentation, the<br />

yeast cells should flocculate, i.e. form aggregates.<br />

Flocculation is dependent on the expression of<br />

a number of surface proteins (flocculins) which<br />

recognize and bind <strong>to</strong> the mannose residues on<br />

mannoproteins located in the outermost wall<br />

layer of other yeast cells (Verstrepen et al., 2003).<br />

These flocculins are probably located in fimbriae,<br />

short hairs (0.5 mm long) on the cell surface<br />

which have been observed by ultrastructural<br />

studies (Day et al., 1975). Numerous environmental<br />

fac<strong>to</strong>rs such as carbon and nitrogen deficiency,<br />

ethanol levels and cell age contribute <strong>to</strong><br />

efficient flocculation, the control of which is one<br />

of the most difficult tasks in brewing. Whether<br />

flocculated yeast accumulates at the <strong>to</strong>p or<br />

bot<strong>to</strong>m of the vessel seems <strong>to</strong> depend on the<br />

flocculins as well as other wall surface properties<br />

(Dengis & Rouxhet, 1997). Attempts are being<br />

made <strong>to</strong> engineer improved brewer’s yeast<br />

strains with respect <strong>to</strong> their flocculation behaviour<br />

(Verstrepen et al., 2003) and also their ability<br />

<strong>to</strong> utilize other carbon sources, including starch<br />

(Hammond, 1995).<br />

Most beer in Europe is brewed in batch fermentations,<br />

but in other countries continuous<br />

cultures following the chemostat principle are<br />

performed. Either way, the ale or lager must be<br />

s<strong>to</strong>red for a while before it is sterilized and<br />

filled in<strong>to</strong> barrels or bottles. In the case of<br />

cask-conditioned Real Ales, the beer is filled<br />

directly in<strong>to</strong> the casks and s<strong>to</strong>red until it is ready<br />

<strong>to</strong> be sold <strong>to</strong> public houses; cask-conditioned ale<br />

is therefore in direct contact with the yeast until<br />

it is served <strong>to</strong> the cus<strong>to</strong>mer. It requires skilled<br />

brewers and publicans <strong>to</strong> keep cask-conditioned<br />

Real Ale, but it does, in the opinion of many,<br />

result in a superior pint.<br />

10.2.10 Wine production<br />

Wine is the fermentation product of starting<br />

materials which already contain high levels of<br />

monosaccharides, i.e. typically fruit juices and<br />

especially the must of grapes. Wine is at least<br />

as ancient as beer and seems <strong>to</strong> have originated<br />

in Transcaucasia and the Near East in the early<br />

Neolithic, around or before 6000 BC (McGovern,<br />

2003). One of the pre-requisites for wine-making<br />

was the invention of pottery, since the fermentation<br />

process requires anaerobic conditions. Wine<br />

has been given a special place in many civilizations<br />

by its association with religious ceremonies,<br />

e.g. in ancient Egypt, Greece, Rome and in<br />

Christianity. In France, Portugal, Luxemburg<br />

and Italy, more than 50 l wine are consumed<br />

per person per annum.<br />

Red wine takes its colour (and high tannin<br />

content) from the skin of the grapes which<br />

are macerated, and the juice (‘must’) is left in<br />

contact with the solid parts for some time. In<br />

contrast, in white wine and rosé wine, the must


276 HEMIASCOMYCETES<br />

is extracted rapidly from the white or red grapes,<br />

respectively. Once the must has been obtained<br />

and filtered, subsequent treatment is similar<br />

for red, rosé and white wines. The must is either<br />

fermented directly, relying on the natural yeast<br />

flora of the grapes (‘spontaneous fermentation’),<br />

or a pure yeast starter culture is added at such<br />

high concentrations that this strain suppresses<br />

the wild yeasts. Spontaneous fermentations still<br />

account for 80% of the worldwide wine production,<br />

and up <strong>to</strong> 100 000 wild yeast cells mostly<br />

not belonging <strong>to</strong> S. cerevisiae may be found on<br />

the surface of one berry or in 1 ml must (Dittrich,<br />

1995). The diversity of yeasts changes rapidly<br />

during the initial stages of the wine fermentation,<br />

with S. cerevisiae displacing the obligately<br />

aerobic species as oxygen becomes depleted. The<br />

<strong>to</strong>tal fruc<strong>to</strong>se and glucose content of musts may<br />

be as high as 150 g l 1 . In principle, fermentation<br />

is completed when no further release of CO 2<br />

occurs, either due <strong>to</strong> exhaustion of sugars or due<br />

<strong>to</strong> ethanol poisoning of the yeast, but in practice<br />

fermentations are often terminated artificially<br />

by addition of sulphite, especially if a sweet wine<br />

is desired.<br />

Fermentation may carry on for up <strong>to</strong> 1 year<br />

with white wine; red wine develops faster but<br />

is often s<strong>to</strong>red in barrels for prolonged periods <strong>to</strong><br />

permit maturation. If red wine is s<strong>to</strong>red in oak<br />

wood, it is called ‘barrique’ wine and it acquires<br />

a characteristic additional flavour. Good concise<br />

accounts of the processes and microbiology of<br />

wine making are those of Dittrich (1995)<br />

and Hartmeier and Reiss (2002).<br />

10.2.11 Production of sake¤<br />

Saké production (for a review, see Oliver, 1991)<br />

involves the conversion of rice starch in<strong>to</strong> monosaccharides<br />

which are then fermented in<strong>to</strong><br />

ethanol. Saké is thus technically a beer rather<br />

than a wine. It has been produced for several<br />

thousand years in China, but its current production<br />

principles based on the synergistic action<br />

of two fungi date back <strong>to</strong> the fifth century AD.<br />

Saké production relies on the degradation of<br />

starch in cleaned and boiled rice by a filamen<strong>to</strong>us<br />

fungus, Aspergillus oryzae, which produces<br />

several different amylases as well as proteases<br />

and other enzymes (see p. 302). Koji, a solid<br />

culture of A. oryzae on steamed rice, is used as<br />

a starter for starch hydrolysis.<br />

Fermentation (moromi) is carried out in<br />

a large volume of water <strong>to</strong> which successive<br />

quantities of boiled rice, koji and the S. cerevisiae<br />

starter culture (mo<strong>to</strong>) are added. Stepwise addition<br />

and a highly ethanol-<strong>to</strong>lerant yeast strain<br />

ensure that saké is the most strongly alcoholic<br />

beverage produced by fermentation without<br />

distillation, containing up <strong>to</strong> 20% (v/v) ethanol.<br />

Fermentation takes about 25 days and is followed<br />

by s<strong>to</strong>rage, maturation and filtration. In order<br />

<strong>to</strong> avoid contamination by lactic acid bacteria,<br />

saké is pasteurized. It is interesting <strong>to</strong> note that<br />

this practice was introduced in the sixteenth<br />

century, 300 years before Pasteur.<br />

10.3 Candida (anamorphic<br />

Saccharomycetales)<br />

Candida is a very large genus of anamorphic<br />

Saccharomycetales, currently comprising some<br />

165 accepted species (Meyer et al., 1998; Kirk et al.,<br />

2001), with new ones being described at a high<br />

frequency. The genus is polyphyletic (Kurtzman<br />

& Robnett, 1998). By far the best-known species<br />

is C. albicans, which is associated with human<br />

disease, and on which we will focus here. A very<br />

similar species, and possibly one which has<br />

been misdiagnosed as C. albicans in the past, is<br />

C. dubliniensis (Martinez et al., 2002). Other species<br />

(C. glabrata, C. inconspicua, C. krusei) may also cause<br />

opportunistic infections of man. In contrast,<br />

Candida utilis (now called Pichia jadinii; see<br />

p. 281) has been used for food and fodder<br />

production for over 80 years, and other Candida<br />

spp. are also suitable for this purpose (Boze et al.,<br />

1995; Scrimshaw & Murray, 1995).<br />

Candida spp. are cosmopolitan and can be<br />

found in many ecological situations (Meyer et al.,<br />

1998), e.g. the surface of fruits and other plant<br />

organs, rotting wood, the soil, sea water, or<br />

associations with mammals and insects (especially<br />

bees). Candida spp. can contaminate grape<br />

musts during the early stages of wine making


CANDIDA (ANAMORPHIC SACCHAROMYCETALES)<br />

277<br />

but are usually displaced by S. cerevisiae later.<br />

Candida albicans is slightly atypical of the genus<br />

in that it does not appear <strong>to</strong> be distributed<br />

widely in the environment and can be considered<br />

a commensal of humans and other warmblooded<br />

animals.<br />

10.3.1 Dimorphism in Candida albicans<br />

Candida albicans can grow as yeast cells, true<br />

septate hyphae, or pseudohyphae which are an<br />

intermediate form between these two extremes<br />

(see Fig. 1.1d). Thick-walled chlamydospores<br />

may be formed by hyphae or pseudohyphae.<br />

This dimorphism or polymorphism has long<br />

been thought <strong>to</strong> represent an important pathogenicity<br />

determinant, pathogenicity commonly<br />

being associated with hyphal growth whereas<br />

yeasts are indicative of saprotrophic commensal<br />

growth. The switch between yeast and hyphal<br />

states is reversible and is determined by an interplay<br />

of several fac<strong>to</strong>rs, e.g. temperature (hyphae<br />

at 37°C, yeasts below), pH (hyphae at neutral<br />

pH, yeasts at acid pH), nutrient abundance<br />

(yeast growth) or deficiency (hyphal growth),<br />

and presence (hyphal growth) or absence<br />

(yeast growth) of blood serum. Thus, conditions<br />

which mimic the bloodstream encourage hyphal<br />

growth, whereas conditions as found on the skin<br />

or in mucosal linings tend <strong>to</strong> promote yeast<br />

growth. Candida albicans is a commensal colonist<br />

of most humans, occasionally causing skin<br />

lesions, but under exceptional circumstances it<br />

turns in<strong>to</strong> a serious pathogen causing deepseated<br />

or systemic mycoses, especially when the<br />

host’s immune system is weakened, e.g. in<br />

AIDS sufferers or patients who have undergone<br />

an organ transplantation. From such infections,<br />

C. albicans is usually recovered in the<br />

hyphal form.<br />

The yeast and hyphal forms differ in many<br />

features which have a bearing on their ability<br />

<strong>to</strong> cause disease (Odds, 1994). For instance,<br />

hyphae are coated with mannoproteins which<br />

adhere strongly <strong>to</strong> mammalian proteins found<br />

in the membranes of cell surfaces. Such adhesive<br />

proteins (adhesins) take the shape of fimbriae<br />

projecting beyond the cell wall (Yu et al., 1994;<br />

Vitkov et al., 2002; see Fig 23.15). Enhanced<br />

adhesion may play a role in pathogenesis,<br />

especially when coupled with the invasive<br />

mode of growth displayed by hyphae (Gow<br />

et al., 1999). Further, hyphae secrete aspartyl proteases<br />

and lipases capable of degrading host<br />

tissue (Hube & Naglik, 2001). Mannoproteins as<br />

well as proteases are potential targets for new<br />

anti-Candida drugs.<br />

Yeast-hyphal dimorphism in C. albicans has<br />

been investigated in some detail. The signalling<br />

chains leading <strong>to</strong> the formation of a hypha are<br />

extremely complex, involving cyclic AMP as well<br />

as mi<strong>to</strong>gen-activated protein kinase (MAP kinase)<br />

pathways. Both are also involved in the switch<br />

from yeast cells <strong>to</strong> pseudohyphal growth in<br />

S. cerevisiae (Brown & Gow, 2002). An extensive<br />

cross-talk between different signalling pathways<br />

is not surprising, since the switch from yeast <strong>to</strong><br />

hypha responds <strong>to</strong> many different environmental<br />

signals which need <strong>to</strong> be integrated. The<br />

control mechanisms determining the switch<br />

from yeast <strong>to</strong> (pseudo)hyphal growth may also<br />

be similar between S. cerevisiae (see Section 10.2.4)<br />

and C. albicans.<br />

10.3.2 Mating and switching in<br />

Candida albicans<br />

Whereas C. albicans is permanently diploid, other<br />

Candida species such as C. glabrata are haploid.<br />

An exclusively diploid vegetative phase is very<br />

unusual among true fungi, although it is found<br />

in Pro<strong>to</strong>myces (Archiascomycetes; see p. 251)<br />

or Xanthophyllomyces (Heterobasidiomycetes; see<br />

Fig. 24.3) and, of course, in the Oomycota<br />

(see Chapter 5). Until recently, C. albicans was<br />

thought <strong>to</strong> reproduce strictly asexually. However,<br />

when the genome sequence of C. albicans became<br />

available and was examined closely, a complete<br />

set of genes relevant <strong>to</strong> mating, homologous<br />

with those known for S. cerevisiae, was detected,<br />

and it was found that the fungus is heterozygous<br />

for the two mating type idiomorphs a and<br />

a, similar <strong>to</strong> the diploid cells of S. cerevisiae but<br />

unable <strong>to</strong> sporulate. The signalling processes<br />

involved in mating are likely <strong>to</strong> be similar<br />

between S. cerevisiae and C. albicans (Bennett &<br />

Johnson, 2005), and conjugation in C. albicans has<br />

now been observed between diploid strains each


278 HEMIASCOMYCETES<br />

Fig10.8 Examples of<br />

white opaque switching in<br />

Candida albicans.(a)Cellsfroma<br />

white colony plated at low density.<br />

A switch has occurred from white<br />

(wh) <strong>to</strong> opaque (op). (b) A white<br />

colony which has been aged on a<br />

plate with limited gas exchange.<br />

This has caused increased rates of<br />

switching from white (wh) <strong>to</strong><br />

opaque (op) at the colony edge.<br />

Original images kindly provided by<br />

D.R. Soll and K. Daniels.<br />

containing only either mating type, although<br />

karyogamy was doubtful and meiosis was not<br />

observed (Magee & Magee, 2000; Lockhart et al.,<br />

2003). If tetraploid strains result from karyogamy<br />

in nature, these may undergo meiosis or random<br />

loss of chromosomes by a parasexual cycle,<br />

i.e. C. albicans may have a cryptic sexual phase<br />

in its life cycle which has eluded mycologists<br />

for over a century (Gow, 2002). Wong et al. (2003)<br />

have suggested that Candida glabrata has a<br />

similarly cryptic sexual cycle.<br />

In contrast <strong>to</strong> S. cerevisiae, there are no silent<br />

additional copies of mating type idiomorphs<br />

in the genome of C. albicans. Before a diploid<br />

strain of C. albicans heterozygous for mating<br />

type idiomorph (i.e. a/a) can mate, it will therefore<br />

have <strong>to</strong> convert <strong>to</strong> a/a or a/a by a mechanism<br />

different from the mating type switch based<br />

on a cassette system as found in S. cerevisiae<br />

(see p. 266).<br />

A remarkable phenomenon that has been<br />

known for some time is the spontaneous and<br />

reversible switching of yeast colony phenotypes<br />

in C. albicans. A given strain can switch its colony<br />

morphology between smooth, wrinkled or starlike,<br />

and white or opaque (Fig. 10.8). The<br />

latter switch is particularly well-characterized<br />

(Slutsky et al., 1987; Soll, 2002) and occurs at<br />

an unusually high frequency (about one colony<br />

in 1000 10 000). The different morphological<br />

appearances of white and opaque colonies are<br />

due <strong>to</strong> differences in size, shape and surface<br />

properties of the yeast cells. Genetically, the<br />

switch is accompanied by the co-ordinated upor<br />

down-regulation of numerous genes, some of<br />

them potentially involved in pathogenesis<br />

(Soll, 2003). Examples are secre<strong>to</strong>ry proteases<br />

or an ABC transporter involved in drug resistance<br />

(see p. 278). Not surprisingly, these two<br />

different colony types have widely differing<br />

pathogenic properties, the white-phase cells<br />

appearing <strong>to</strong> be better adapted <strong>to</strong> colonization<br />

of internal organs and opaque-phase cells superior<br />

in colonizing external skin regions.<br />

An interesting link between mating and<br />

switching is that the latter is suppressed in<br />

strains heterozygous for the mating type idiomorphs<br />

a and a. This may be mediated by the<br />

regula<strong>to</strong>ry heterodimer presumably formed by<br />

protein products of the a and a idiomorphs.<br />

Another noteworthy observation involves sexual<br />

reproduction: opaque cells mate about 10 6<br />

times more efficiently than white cells (Miller &<br />

Johnson, 2002). Mating competence in C. albicans<br />

is therefore regulated at two levels, namely the<br />

requirement for a given cell <strong>to</strong> be homozygous<br />

for mating type (either a or a) and <strong>to</strong> be in the<br />

opaque state (Soll et al., 2003). It is furthermore<br />

possible that some of the phenotypes characteristic<br />

of opaque cells are required for mating, in<br />

addition <strong>to</strong> or instead of being pathogenicity<br />

fac<strong>to</strong>rs. In this context, it is of interest that<br />

there is a mass switch from opaque <strong>to</strong> white<br />

at 37°C, and that mating between opaque cells<br />

of opposite mating type is strongly stimulated<br />

on skin surfaces which have a lower temperature<br />

(Lachke et al., 2003). Clearly, the ability of<br />

C. albicans <strong>to</strong> adapt <strong>to</strong> different situations by<br />

changing between several pre-programmed cell<br />

types, e.g. the mating-competent opaque and the<br />

invasive white cells, contributes <strong>to</strong> the success


CANDIDA (ANAMORPHIC SACCHAROMYCETALES)<br />

279<br />

of this organism as a pathogen (Staib et al., 2001;<br />

Bennett & Johnson, 2005).<br />

10.3.3 Treatment of candidiasis and<br />

resistance mechanisms<br />

Candida infections often occur following treatment<br />

with antibacterial antibiotics which also<br />

kill the benign bacteria which compete against<br />

Candida. Such superficial infections are especially<br />

common in mucosal linings of the mouth cavity,<br />

vagina or on the skin. They are collectively called<br />

‘thrush’. Infections of the oesophagus occur in<br />

patients with weakened immune system and are<br />

considered an AIDS-defining illness. The mucous<br />

membranes and skin are usually effective as<br />

primary barriers against infection, and Candida<br />

cells within the human body are vigorously<br />

attacked by the immune system (Murphy, 1996;<br />

Magliani et al., 2005). If all these barriers are<br />

broken or weakened, deep invasive candidiasis<br />

can occur. Yeast cells (conidia) can be disseminated<br />

in the blood stream, and individual organs<br />

can become colonized by hyphae. Contaminated<br />

catheters are also an important entry point<br />

for Candida. An extended account of candidiasis<br />

in all its forms has been given by Kwon-Chung<br />

and Bennett (1992).<br />

Generally, treatment of Candida infections<br />

is difficult because of the relative genetic similarity<br />

between Candida and humans, which<br />

greatly reduces the range of available targets<br />

as compared <strong>to</strong> the treatment of bacterial<br />

infections. Consequently, certain anti-Candida<br />

drugs have severe side effects. None the less,<br />

drugs belonging <strong>to</strong> several different classes<br />

are in current clinical use, as reviewed by<br />

Georgopapadakou (1998), Cowen et al. (2002)<br />

and Sanglard and Bille (2002). Lucid accounts of<br />

the fascinating array of mechanisms by which<br />

Candida achieves resistance against the various<br />

drugs are those by Ghannoum and Rice (1999),<br />

Sanglard (2002) and Akins (2005).<br />

Many drugs target ergosterol, a fungal<br />

membrane sterol which is not found in animals.<br />

Amphotericin B (Fig. 10.9a) or nystatin are<br />

polyene antibiotics which associate with ergosterol<br />

in the membranes of Candida, forming<br />

pores in the plasma membrane and thereby rendering<br />

it leaky. Amphotericin B has severe<br />

side effects but has <strong>to</strong> be used especially against<br />

deep-seated infections (Lemke et al., 2005).<br />

Resistance is usually based on the replacement<br />

of ergosterol by a precursor molecule, or a general<br />

reduction of the sterol content in the plasma<br />

membrane.<br />

Fig10.9 The most important anti-Candida drugs in current use. (a) The polyene compound amphotericin B. (b) The triazole<br />

compound fluconazole. (c) The allylamine terbinafine. (d) The fluoropyrimidine compound 5-fluorocy<strong>to</strong>sine. (e) The echinocandin<br />

caspofungin.


280 HEMIASCOMYCETES<br />

The azole-type fungicides inhibit the enzyme<br />

lanosterol demethylase which is involved in<br />

ergosterol biosynthesis (see Fig. 13.16). A muchused<br />

example is fluconazole (Fig. 10.9b) which is<br />

free from severe side effects. There are several<br />

resistance mechanisms in C. albicans. The most<br />

common, found in 85% of all resistant isolates<br />

(Perea et al., 2001), is based on active exclusion of<br />

the drug by means of ABC (ATP binding cassette)<br />

transporters or similar mechanisms. These are<br />

plasma membrane proteins with numerous<br />

(usually 12) transmembrane domains and two<br />

cy<strong>to</strong>plasmic ATP-binding domains. Alternating<br />

binding and hydrolysis of ATP changes the conformation<br />

of these proteins, enabling them <strong>to</strong><br />

open and close membrane pores. ABC transporters<br />

are often capable of transporting a group<br />

of different metabolites, thereby producing<br />

cross-resistance. The natural role of ABC transporters<br />

probably lies in the exclusion of endogenous<br />

antibiotics, <strong>to</strong>xins or other substances, e.g.<br />

mating fac<strong>to</strong>rs in S. cerevisiae. The second most<br />

important type of resistance against azole-type<br />

drugs (65% of all isolates) is a mutation of the<br />

cellular target, i.e. the azole binding site on the<br />

enzyme lanosterol demethylase. The fact that<br />

the occurrences of these two types of resistance<br />

add up <strong>to</strong> more than 100% indicates that many<br />

clinical C. albicans isolates (about 75%) possess<br />

both resistance mechanisms. A third resistance<br />

mechanism against azoles is the overexpression<br />

of the gene ERG11 encoding lanosterol demethylase,<br />

which occurs in about 35% of resistant<br />

isolates (Perea et al., 2001).<br />

A third group of compounds, the allylamines<br />

(e.g. terbinafine; Fig. 10.9c) act against a different<br />

enzyme involved in ergosterol biosynthesis, squalene<br />

epoxidase (see Fig. 13.16). Terbinafine is<br />

not used extensively on its own, but it is useful<br />

in combination with other drugs in order <strong>to</strong><br />

treat infections by resistant Candida strains.<br />

Another target, nucleic acid biosynthesis,<br />

is attacked by 5-fluorocy<strong>to</strong>sine (Fig. 10.9d).<br />

Following uptake, it is deaminated <strong>to</strong> 5-fluorouracil<br />

and converted <strong>to</strong> 5-fluoro-UTP or a corresponding<br />

deoxynucleotide, which inhibit RNA<br />

and DNA biosynthesis, respectively. Resistance is<br />

associated with a reduced capacity of the fungus<br />

<strong>to</strong> metabolize 5-fluorocy<strong>to</strong>sine. Mammalian cells<br />

do not efficiently metabolize this drug but intestinal<br />

bacteria can, which precludes the oral use<br />

of this antibiotic.<br />

A recently described group are the echinocandins<br />

which inhibit b-(1,3)-glucan synthesis.<br />

There is preliminary evidence that echinocandins<br />

such as caspofungin (Fig. 10.9e) do not<br />

act directly on b-(1,3)-synthase but in an indirect<br />

manner by interfering with upstream regula<strong>to</strong>ry<br />

proteins (Edlind & Katiyar, 2004). No drugs<br />

against mannoproteins or aspartic proteases are<br />

as yet commercially available, although treatment<br />

against HIV uses protease inhibi<strong>to</strong>rs which<br />

also affect C. albicans (Dupont et al., 2000).<br />

Research efforts in<strong>to</strong> new anti-Candida drugs are<br />

intensive, given that the incidence of Candida<br />

infections is strongly on the increase, few substances<br />

are currently available, and resistance of<br />

Candida against them is becoming a problem.<br />

10.3.4 Ecology and drug resistance of<br />

Candida albicans<br />

Numerous investigations of the distribution of<br />

Candida spp. on their hosts have been carried out.<br />

Generally, C. albicans is by far the most frequent<br />

species, followed by C. parapsilosis. Other species<br />

such as C. glabrata, C. krusei and C. tropicalis<br />

are very much less frequent. Within the species<br />

C. albicans, many different strains exist, and their<br />

colonization pattern has been followed on the<br />

same human host over time (Xu et al., 1999;<br />

Kam & Xu, 2002). Each human being can be<br />

colonized by a diversity of strains. Displacement<br />

of one strain by another is possible, as is the<br />

transfer of strains between humans. Appropriate<br />

analyses of allelic distributions have shown that<br />

the mode of genetic inheritance is predominantly<br />

clonal, i.e. sexual reproduction and<br />

the exchange of genetic material between different<br />

Candida strains do not seem <strong>to</strong> play an<br />

important role (Lott et al., 1999). There are no<br />

significant differences in the Candida populations<br />

between healthy individuals and AIDS<br />

patients unless, of course, the population<br />

dynamics are shaken up by anti-Candida drug<br />

treatments. In the course of a prolonged treatment<br />

of patients against oral candidiasis,<br />

Martinez et al. (2002) reported the displacement


GALACTOMYCES (DIPODASCACEAE)<br />

281<br />

of the initially predominant, fluconazolesensitive<br />

C. albicans flora by fluconazoleresistant<br />

C. dubliniensis strains especially in<br />

those patients where C. albicans had failed <strong>to</strong><br />

develop resistance. Resistance may arise spontaneously<br />

after prolonged treatment, and the<br />

spread of resistant clones in hospitals may not<br />

be as important with Candida as, for example,<br />

with multiple drug-resistant bacteria (Taylor<br />

et al., 2003).<br />

processes of proteins of pharmacological interest<br />

(Daly & Hearn, 2005).<br />

10.5 Galac<strong>to</strong>myces (Dipodascaceae)<br />

The genus Galac<strong>to</strong>myces (formerly called Endomyces)<br />

is characterized by true hyphae which<br />

10.4 Pichia (Saccharomycetaceae)<br />

The genus Pichia contains 94 species (Kurtzman,<br />

1998; Kirk et al., 2001) and is characterized by<br />

budding cells, with only a few species also<br />

producing arthroconidia, pseudohyphae and<br />

hyphae. Sexual reproduction is by ascospores<br />

(1 4 per ascus) which are often hat-shaped<br />

(galeate). Molecular characterization of the<br />

genus is still in progress (Suzuki & Nakase, 1999)<br />

and will undoubtedly lead <strong>to</strong> rearrangements<br />

in future.<br />

Pichia is cosmopolitan and ubiqui<strong>to</strong>us. A<br />

surprising number of species has been isolated<br />

from the frass of wood-attacking beetles<br />

(Kurtzman, 1998); others grow on the exudates<br />

(slime fluxes) of trees or on decaying cacti, or<br />

they occur as contaminants of industrial fermentations.<br />

Two species are of particular biotechnological<br />

interest: Pichia jadinii (anamorph Candida<br />

utilis), formerly called Torula yeast, has been<br />

developed since World War I as a food yeast<br />

for single-cell protein. It can utilize the pen<strong>to</strong>ses<br />

of pulping-waste liquors from the paper industry<br />

and is also grown on other biological wastes<br />

(Boze et al., 1995).<br />

Pichia pas<strong>to</strong>ris is interesting for a different<br />

reason; it can utilize methanol by expressing<br />

and secreting large quantities of alcohol oxidase.<br />

Since the protein glycosylation chains of<br />

P. pas<strong>to</strong>ris are similar <strong>to</strong> those of humans, and<br />

because the products of heterologous genes are<br />

secreted efficiently, P. pas<strong>to</strong>ris has advantages<br />

over S. cerevisiae in the industrial production<br />

Fig10.10 Galac<strong>to</strong>myces candidus. (a) Vegetative hyphal apex.<br />

The two lateral branches near the base are developing<br />

conidiophores. (b) Conidiophore showing the development<br />

and separation of arthroconidia. (c) Gametangia developing as<br />

lateral bulges of the hyphae on either side of a septum.<br />

(d) Fusion of gametangia <strong>to</strong> form asci. In one ascus, a single<br />

ascospore is differentiated. (e) Mature asci, each containing<br />

a single ascospore.


282 HEMIASCOMYCETES<br />

form septa and quickly fragment in<strong>to</strong> arthrospores<br />

(Fig. 10.10), giving the colonies a creamy<br />

appearance. The septa have micropores, like<br />

those of true hyphae of C. albicans. Six species<br />

are now known; by far the most important is<br />

G. candidus (anamorph Geotrichum candidum),<br />

formerly known as Galac<strong>to</strong>myces geotrichum (de<br />

Hoog & Smith, 2004). It is a ubiqui<strong>to</strong>us mould<br />

which is common in soil, dairy products, sewage<br />

and other substrata. It is also thought <strong>to</strong> be a<br />

common constituent of the skin and gut flora<br />

of humans and animals, although reports of<br />

it being a human pathogen have generally<br />

remained unsubstantiated (Kwon-Chung &<br />

Bennett, 1992). In addition, G. candidus is a wellknown<br />

cause of post-harvest rot in ripe fruits<br />

and vegetables especially when these are kept<br />

in plastic bags. Infected plant organs become<br />

soft and eventually, upon puncturing, exude<br />

a creamy mass of decaying tissue which has<br />

a sour smell; hence the name ‘sour rot’ (Agrios,<br />

2005).<br />

The fungus grows readily in culture, forming<br />

broad hyphae with finer lateral branches. The<br />

vegetative cells contain 1 4 nuclei. Branching<br />

is of two kinds, pseudodicho<strong>to</strong>mous near the<br />

apex, and lateral immediately behind a septum.<br />

It is from such lateral branches that conidia<br />

develop (Figs. 10.10a,b). Conidiophores are difficult<br />

<strong>to</strong> differentiate from vegetative hyphae.<br />

Prior <strong>to</strong> conidium formation, apical growth of<br />

a hypha ceases, then septa are laid down in the<br />

tip region. The septa are two-ply, and separation<br />

of the two layers making up the septum<br />

leads <strong>to</strong> the disarticulation of the terminal part<br />

of a hypha in<strong>to</strong> cylindrical segments termed<br />

arthrospores or arthroconidia (Cole & Kendrick,<br />

1969b). Conidia of other Galac<strong>to</strong>myces spp. are<br />

virtually indistinguishable from those of<br />

G. candidus.<br />

Galac<strong>to</strong>myces candidus may be homo- or heterothallic,<br />

but the sexual state is not frequently<br />

seen. After mating, fertile hyphae are produced<br />

and gametangia arise in pairs on either side of<br />

a septum, in the broad main hyphae or short<br />

side branches (Figs. 10.10c e). Fusion of the<br />

gametangia gives rise <strong>to</strong> a globose fusion cell<br />

which becomes transformed directly in<strong>to</strong> an<br />

ascus. The ascus contains only a single ascospore<br />

which has two wall layers, a smooth inner layer<br />

and a furrowed outer layer. Each ascospore<br />

contains 1 2 nuclei. Whether and when meiosis<br />

occurs is not yet known.<br />

The Geotrichum arthroconidial state is found<br />

also in the only other genus of the Dipodascaceae,<br />

Dipodascus, which produces multispored<br />

asci with 4 128 spores. A superficially similar<br />

state, Saprochaete, is formed by a genus of phylogenetically<br />

unrelated fungi now called Magnusiomyces.<br />

Species descriptions and a key of<br />

Galac<strong>to</strong>myces, Dipodascus and Magnusiomyces have<br />

been provided by de Hoog and Smith (2004).<br />

10.6 Saccharomycopsis<br />

(Saccharomycopsidaceae)<br />

Saccharomycopsis (formerly Endomycopsis) is a mycelial<br />

yeast which reproduces by buds (blas<strong>to</strong>spores<br />

or yeast cells) and also forms asci parthenogenetically<br />

or following isogamous fusion. About<br />

10 species are known (Kurtzman & Smith, 1998;<br />

Barnett et al., 2000). Saccharomycopsis fibuligera<br />

grows in flour, bread, macaroni and other<br />

starchy substrates, and produces a complex of<br />

numerous active extracellular amylases, an<br />

unusual property in yeasts (Hostinová, 2002).<br />

This has been used <strong>to</strong> develop S. fibuligera as<br />

a food yeast for cattle feed which can be grown<br />

on pota<strong>to</strong> starch processing wastes (Jarl, 1969).<br />

This species is also used extensively for starch<br />

hydrolysis by starter cultures in Far Eastern<br />

fermented food (Beuchat, 1995).<br />

In culture, S. fibuligera may form budding<br />

yeast cells and branched septate hyphae which<br />

produce blas<strong>to</strong>spores laterally and terminally<br />

(Fig. 10.11). Arthrospore formation has also<br />

been demonstrated. Ascus formation in this<br />

homothallic species can be induced by growing<br />

the yeast for a few days on malt extract agar<br />

and transferring it <strong>to</strong> distilled water. The asci<br />

are mostly four-spored, and the spores are<br />

hat-shaped (Fig. 10.11d), having a flange-like<br />

extension of the wall.


SACCHAROMYCOPSIS (SACCHAROMYCOPSIDACEAE)<br />

283<br />

Fig10.11 Saccharomycopsis fibuligera. (a,b) Mycelium from three-day-old culture showing blas<strong>to</strong>spore formation. (c) Blas<strong>to</strong>spores<br />

germinating by germ tube, or budding <strong>to</strong> form a further blas<strong>to</strong>spore. (d) Ayoung ascus and two mature asci containing four<br />

hat-shaped ascospores. (e) Germinating ascospore. (a,c e) <strong>to</strong> same scale.<br />

Fig10.12 Scanning electron micrographs of Saccharomycopsis javanensis preying upon Saccharomyces cerevisiae. (a) Points of<br />

penetration (arrowheads) at an early stage. Also note the bud-scars on an older S. cerevisiae cell in the centre of the picture.<br />

(b) Collapsed cells of penetrated prey (arrows) at a later stage.Original images kindly provided by M.-A. Lachance. Reprinted<br />

from Lachance et al. (2000) by copyright permission of the National Research Council of Canada.


284 HEMIASCOMYCETES<br />

Fig10.13 Eremothecium coryli.(a)Vegetative<br />

growth as a mass of hyphae and yeast cells. (b)<br />

Ascus containing eight ascospores. Both<br />

images <strong>to</strong> same scale.<br />

Most members of the genus Saccharomycopsis<br />

have been observed <strong>to</strong> behave as predacious<br />

yeasts on leaf surfaces, i.e. they attack, penetrate<br />

and digest the cells of other yeasts (Fig. 10.12).<br />

This phenomenon is different from that of killer<br />

yeasts because no <strong>to</strong>xins appear <strong>to</strong> be involved.<br />

Instead, penetration and killing is brought<br />

about mainly by cell wall-degrading enzymes,<br />

notably b-(1,3)-glucanase (Lachance et al., 2000).<br />

It is interesting <strong>to</strong> note that all those yeasts<br />

capable of preying on others are incapable<br />

of utilizing sulphate as a source of sulphur,<br />

although not all yeasts deficient in sulphate<br />

transport are predacious. Predation can be cannibalistic,<br />

but many other yeasts belonging <strong>to</strong><br />

the Asco- and Basidiomycota are also attacked.<br />

Clearly, the phylloplane is a highly competitive<br />

environment.<br />

10.7 Eremothecium<br />

(Eremotheciaceae)<br />

This genus contains five species which were<br />

formerly classified in different genera but have<br />

now been shown <strong>to</strong> be closely related by DNA<br />

sequence analyses (Kurtzman, 1995; de Hoog<br />

et al., 1998). Pseudohyphae and true hyphae are<br />

present in culture, and vegetative reproduction<br />

is often by budding yeast cells. Asci are formed<br />

terminally or in intercalary positions, and<br />

they contain 8 32 needle-shaped ascospores<br />

(Fig. 10.13).<br />

Eremothecium spp. are the only important<br />

plant pathogens among the Hemiascomycetes<br />

and can infect numerous plant species, causing<br />

damage especially on cot<strong>to</strong>n (Gossypium spp.).<br />

They are transmitted by hemipteran insects<br />

which may harbour inoculum in their stylet<br />

pouches. The route of entry in<strong>to</strong> the plant<br />

is often via the stigma of the flower (Batra,<br />

1973). Eremothecium (formerly Nema<strong>to</strong>spora) coryli<br />

(Fig. 10.13) causes a disease called stigma<strong>to</strong>mycosis<br />

on a wide range of plants, including hazel<br />

(Corylus).<br />

In biotechnology, E. ashbyi and E. (Ashbya)<br />

gossypii are used in fermentations for the


11<br />

Plec<strong>to</strong>mycetes<br />

11.1 <strong>Introduction</strong><br />

The class Plec<strong>to</strong>mycetes originally contained all<br />

ascomycetes which produce their asci within a<br />

cleis<strong>to</strong>thecium, i.e. a ‘closed case’. DNA sequence<br />

comparisons have revealed that this character<br />

was a fairly good one because, with the major<br />

exception of the powdery mildews (Erysiphales;<br />

see Chapter 13) and few scattered examples in<br />

the Pyrenomycetes (Chapter 12) and Helotiales<br />

(Chapter 15), most cleis<strong>to</strong>thecium-forming fungi<br />

and the anamorphs associated with them have<br />

been found <strong>to</strong> be monophyletic (Berbee & Taylor,<br />

1992a; Geiser & LoBuglio, 2001). As they stand<br />

now, the Plec<strong>to</strong>mycetes can be defined by the<br />

following set of characters (Alexopoulos et al.,<br />

1996; Geiser & LoBuglio, 2001).<br />

(1) A cleis<strong>to</strong>thecium or gymnothecium is<br />

usually present; a cleis<strong>to</strong>thecium proper has a<br />

thick and continuous (pseudoparenchyma<strong>to</strong>us)<br />

wall, whereas in the gymnothecium the wall<br />

consists of an open cage-like construction<br />

of hyphae, the reticuloperidium (Greif &<br />

Currah, 2003). Naked asci are produced only in<br />

rare cases.<br />

(2) Ascogenous hyphae are usually not<br />

conspicuous.<br />

(3) Asci are scattered throughout the cleis<strong>to</strong>thecium,<br />

not produced by a fertile layer<br />

(hymenium).<br />

(4) Asci are mostly globose and thin-walled,<br />

and the ascospores are released passively after<br />

disintegration of the ascus wall, not by active<br />

discharge.<br />

(5) Ascospores are small, unicellular and<br />

usually spherical or ovoid.<br />

(6) Conidia are commonly produced from<br />

phialides (in Eurotiales) or as arthroconidia,<br />

which are typically formed as chains of conidia<br />

alternating with sterile cells. An arthroconidium<br />

becomes released when the neighbouring cells<br />

disintegrate. This rhexolytic secession is typical<br />

of the microconidia of Onygenales and<br />

Ascosphaerales. The alternative is schizolytic<br />

secession in which adjacent cells separate when<br />

the septum joining them splits in<strong>to</strong> two (see<br />

Figs. 8.9, 10.10), but this is not found in the<br />

Plec<strong>to</strong>mycetes. However, terminal thick-walled<br />

chlamydospores and multicellular blastic macroconidia<br />

may be produced by some Plec<strong>to</strong>mycetes.<br />

Plec<strong>to</strong>mycetes are predominantly saprotrophic<br />

fungi associated with the soil. Many have<br />

a capacity <strong>to</strong> degrade complex biopolymers, e.g.<br />

starch and cellulose, while others degrade<br />

proteins such as keratin which makes up hair,<br />

horn and feathers. If proteolytic fungi can grow<br />

at 37°C, they are potentially pathogenic <strong>to</strong><br />

mammals, and some of them are indeed among<br />

the most dangerous fungal pathogens of man.<br />

Many other Plec<strong>to</strong>mycetes produce important<br />

secondary metabolites, e.g. antibiotics and<br />

myco<strong>to</strong>xins.<br />

Several taxonomic arrangements have been<br />

proposed for the Plec<strong>to</strong>mycetes (e.g. Kirk et al.,<br />

2001; Eriksson et al., 2003), but we have chosen<br />

that of Geiser and LoBuglio (2001) because of its<br />

clarity. This divides the Plec<strong>to</strong>mycetes in<strong>to</strong> three<br />

orders, the Ascosphaerales, Onygenales, and<br />

Eurotiales (Table 11.1). We will consider


286 PLECTOMYCETES<br />

Table 11.1. Classification of Plec<strong>to</strong>mycetes following Geiser and LoBuglio (2001) and Kirk et al. (2001).<br />

Order Family No. oftaxa Examples of<br />

teleomorphs<br />

Examples of<br />

anamorphs<br />

Ascosphaerales<br />

(see Section11.2)<br />

Onygenales<br />

(see Section11.3)<br />

Eurotiales<br />

(see Section11.4)<br />

Ascosphaeraceae 3 gen.,13 spp. Ascosphaera (mostly unknown)<br />

Eremascaceae 1gen., 2 spp. Eremascus (unknown)<br />

Onygenaceae<br />

(see p. 290)<br />

Arthrodermataceae<br />

(see p. 293)<br />

Gymnoascaceae<br />

(see p. 295)<br />

Myxotrichaceae<br />

(see p. 295)<br />

Trichocomaceae<br />

(see p. 297)<br />

22 gen.,57 spp. Ajellomyces,<br />

Auxarthron,<br />

Amauroascus,<br />

Onygena<br />

Malbranchea,<br />

Chrysosporium,<br />

Coccidioides,<br />

His<strong>to</strong>plasma,<br />

Paracoccidioides,<br />

Blas<strong>to</strong>myces<br />

2gen.,48spp.<br />

Ctenomyces, Chrysosporium,<br />

Arthroderma Microsporum,<br />

Epidermophy<strong>to</strong>n,<br />

Trichophy<strong>to</strong>n<br />

10 gen., 23 spp. Gymnoascus (mostly unknown)<br />

4gen.,12spp. Myxotrichum Geomyces,<br />

Malbranchea,<br />

Oidiodendron<br />

20 gen., 4500 spp. Byssochlamys,<br />

Emericella,<br />

Eupenicillium,<br />

Eurotium,<br />

Talaromyces<br />

Aspergillus,<br />

Paecilomyces,<br />

Penicillium<br />

Monascaceae 2 gen.,7 spp. Monascus Basipe<strong>to</strong>spora<br />

Elaphomycetaceae 1gen.,20 spp. Elaphomyces (unknown)<br />

(see p. 313)<br />

representatives from all three orders, with a<br />

particular emphasis on the Eurotiales which<br />

contain the important anamorphic genera<br />

Aspergillus and Penicillium.<br />

11.2 Ascosphaerales<br />

This small order currently comprises the 4<br />

teleomorphic genera Arrhenosphaera (1 species),<br />

Ascosphaera (11 species), Bettsia (1 species) and<br />

Eremascus (2 species). The first three genera are<br />

associated with beehives whereas Eremascus is<br />

a food-spoilage fungus. All genera can grow on,<br />

and sometimes require, substrates rich in sugar<br />

or salt. Some species are truly xerophilic, i.e. they<br />

can grow at water activities (a W ) lower than 0.85,<br />

which is equivalent <strong>to</strong> a solution containing<br />

60% glucose. Ascosphaerales are atypical of<br />

the Plec<strong>to</strong>mycetes because they do not produce<br />

true cleis<strong>to</strong>thecia, but DNA-based phylogenetic<br />

studies have shown that they belong here (Berbee<br />

et al., 1995).<br />

11.2.1 Eremascus<br />

In mycology as in many other areas of biology,<br />

it is virtually impossible <strong>to</strong> establish a rule without<br />

having <strong>to</strong> qualify it almost immediately<br />

by giving exceptions and modifications <strong>to</strong> it.<br />

Eremascus is a member of the Euascomycete clade<br />

with free asci which are not organized in<strong>to</strong>


ASCOSPHAERALES<br />

287<br />

ascocarps, and it is thus an exception <strong>to</strong> the<br />

generalization which places such fungi in the<br />

Archiascomycetes (Chapter 9) or Hemiascomycetes<br />

(see Eremothecium, Galac<strong>to</strong>myces, Saccharomycopsis;<br />

Chapter 10). Support for the inclusion of<br />

Eremascus in the Euascomycetes comes not only<br />

from molecular data (Berbee & Taylor, 1992a;<br />

Anderson et al., 1998) but also from the presence<br />

of typical Euascomycete septa with one central<br />

pore and associated Woronin bodies (Kreger-van<br />

Rij et al., 1974). Further, the arthroconidia are<br />

delimited by a double-septum (Harrold, 1950)<br />

and are released by rhexolytic secession, which is<br />

typical of certain Plec<strong>to</strong>mycetes (see Fig. 11.3f).<br />

Two species are known, E. albus and E. fertilis<br />

(Fig. 11.1). Both are associated with sugary substrates<br />

such as mouldy jam, but several collections<br />

of E. albus have been made from powdered<br />

mustard. Harrold (1950) has shown that both<br />

fungi grow best on media with a high sugar<br />

content (e.g. 40% sucrose), but do not grow well<br />

in a water-saturated atmosphere. The mature<br />

mycelium consists of uninucleate segments.<br />

Both species are homothallic. On either side of<br />

a septum, short gametangial branches arise<br />

which are swollen at their tips and, in the case<br />

of E. albus, coil around each other. The gametangial<br />

tips of E. albus are usually uninucleate and,<br />

following breakdown of the wall separating the<br />

tips of adjacent gametangia, nuclear fusion<br />

occurs. This is followed by meiosis and mi<strong>to</strong>sis<br />

so that eight nuclei result, each one being<br />

surrounded by cy<strong>to</strong>plasm <strong>to</strong> form a uninucleate<br />

ascospore (Fig. 11.1m). The ascospores are<br />

dispersed passively following breakdown of the<br />

ascus wall. On germination a multinucleate germ<br />

tube emerges, but the uninucleate condition is<br />

soon established by the formation of septa.<br />

11.2.2 Ascosphaera<br />

Ascosphaera spp. are associated with bees and<br />

related insects, growing saprotrophically in their<br />

nests on the gathered pollen and nectar. They can<br />

be maintained in pure culture but commonly<br />

Fig11.1 Eremascus.(a d) Eremascus fertilis, stages in the development of asci. (e g) Eremascus albus, stages in the development<br />

of asci. Note the coiling of the gametangia and the globose ascospores of E. albus.(h m) Eremascus albus, nuclear behaviour<br />

during ascus formation (after Harrold,1950). (h) Uninucleate gametangia. (i) Plasmogamy and karyogamy. (k m) Nuclear divisions<br />

preceding ascospore formation.


288 PLECTOMYCETES<br />

Fig11.2 Ascosphaera apis. (a) Dead infected bee larvae. (b) Young sporocyst. (c) Several intact mature sporocysts. (d) Ruptured<br />

sporocyst with released ascospore balls.<br />

require up <strong>to</strong> 40% glucose in the medium. On<br />

lower-strength agar media they may still grow but<br />

often fail <strong>to</strong> produce ascospores. Ascosphaera apis is<br />

a pathogenic species causing ‘chalk brood’ disease<br />

of honey bees (Apis mellifera). Dead infected larvae<br />

appear white and hard like chalk (Fig. 11.2a), and<br />

black spore balls may break through the integuments<br />

(Skou, 1972, 1975, 1988). The disease occurs<br />

as an epidemic in some years and can seriously<br />

weaken bee colonies, especially if accompanied by<br />

pests such as Varroa mite infestations.<br />

Some species are homothallic but A. apis is<br />

heterothallic. Ascospores are produced in a<br />

unique structure termed a sporocyst (Skou,<br />

1982). The ‘female’ colony produces an ascogonium<br />

terminating in a trichogyne, and plasmogamy<br />

occurs between the trichogyne and an<br />

undifferentiated hypha of the opposite mating<br />

type. Following plasmogamy, the trichogyne<br />

grows backwards in<strong>to</strong> the ascogonium, the wall<br />

of which swells greatly <strong>to</strong> form the sporocyst<br />

(Fig. 11.2b; Spil<strong>to</strong>ir, 1955). When it is mature,<br />

the sporocyst acquires a brown pigmentation<br />

(Figs. 11.2c,d). Within the sporocyst, a system of<br />

binucleate cells with croziers forms eight-spored<br />

asci in clusters. The ascus walls are evanescent,<br />

and the ascospores from the asci of any one<br />

cluster stick <strong>to</strong>gether as spore balls which are<br />

released when the sporocyst wall breaks<br />

(Figs. 11.2c,d). The sporocyst is not homologous<br />

with a cleis<strong>to</strong>thecium because it arises<br />

from a single cell which enlarges prior <strong>to</strong> formation<br />

of the asci, whereas a cleis<strong>to</strong>thecium is<br />

multicellular and grows around the developing<br />

asci.<br />

In order <strong>to</strong> produce ascospores on infected<br />

bee larvae, A. apis requires a slight reduction of<br />

temperature (normally around 33 36°C in intact<br />

hives) <strong>to</strong> about 30°C. Infections by A. apis are<br />

usually most severe in cool weather, especially<br />

in spring. Interestingly, bee colonies have been<br />

found <strong>to</strong> respond <strong>to</strong> A. apis infections by<br />

elevating their temperature, and this so-called<br />

‘behavioural fever’ may retard the outbreak of<br />

the disease (Starks et al., 2000). A further way for<br />

bees <strong>to</strong> control the disease is hygiene, i.e. they<br />

uncap brood cells and remove dead larvae before<br />

A. apis can sporulate on them. Bees can be bred<br />

for hygiene, and the basis of this is thought<br />

<strong>to</strong> be an enhanced sensitivity <strong>to</strong> the odour of


ONYGENALES<br />

289<br />

infected larvae rather than hygienic behaviour<br />

per se which is instinctive (Masterman<br />

et al., 2001). Larvae become infected by A. apis<br />

by ingestion. Many types of commercially available<br />

honey contain viable spores of A. apis<br />

(Anderson et al., 1997).<br />

11.3 Onygenales<br />

This order of the Plec<strong>to</strong>mycetes is of utmost<br />

significance <strong>to</strong> medical mycologists because it<br />

contains most of the true human pathogens, i.e.<br />

fungi able <strong>to</strong> cause disease in otherwise healthy<br />

and immunocompetent individuals. Some taxonomic<br />

confusion has arisen because many of the<br />

serious pathogens have been known for a long<br />

time only in their anamorphic form and continue<br />

<strong>to</strong> be called by their anamorphic names. The<br />

current Dictionary of the <strong>Fungi</strong> (Kirk et al., 2001)<br />

recognizes three families Arthrodermataceae,<br />

Gymnoascaceae and Onygenaceae but it<br />

excludes the family Myxotrichaceae which is of<br />

uncertain placement (incertae sedis), possibly<br />

belonging <strong>to</strong> the Helotiales (Tsuneda & Currah,<br />

2004). Since members of this last family have<br />

many features in common with the other three,<br />

we will consider them briefly here. Including the<br />

Myxotrichaceae, there are some 120 species in the<br />

Onygenales.<br />

Defining features of the Onygenales are that<br />

their ascoma consists of loosely interwoven and<br />

often thick-walled hyphae which sometimes bear<br />

complex and species-characteristic appendages<br />

(Figs. 11.3a, 11.5, 11.8, 11.9). Such a cage-like<br />

ascoma is termed gymnothecium, and the<br />

meshwork of hyphae making up the basket<br />

(peridium) is called reticuloperidium. Greif<br />

and Currah (2003) have shown that the reticuloperidium<br />

can be pierced by the stiff hairs of<br />

arthropods such as flies, and gymnothecial<br />

appendages may also be caught by the limbs<br />

of flies during grooming. Movements by the<br />

animals shake the ascospores out of the<br />

Fig11.3 Onygenaceae. (a) Quarter-segment of a gymnothecium of Ajellomyces capsulatus showing coiled appendages. (b) Ascospore<br />

of A. capsulatus. (c) Tuberculate macroconidia of His<strong>to</strong>plasma capsulatum. (d) Microconidia of H. capsulatum. (e) The‘pilot wheel’ stage<br />

of Paracoccidioides brasiliensis. One giant yeast cell is producing several buds. (f) Malbranchea-type arthroconidia.The conidia are<br />

released by rhexolytic secession, i.e. conidia are spaced apart by sterile cells which eventually disintegrate. (a,c e) <strong>to</strong> same scale.<br />

Redrawn and modified from de Hoog et al. (2000a), with kind permission of Centraalbureau voor Schimmelcultures.


290 PLECTOMYCETES<br />

Fig11.4 Onygena. (a) Stalked gymnothecial stromata of<br />

O. equina on a cast sheep’s horn. (b) Ascogenous hyphae and<br />

asci of O. corvina.<br />

gymnothecium and distribute them. Thus, the<br />

gymnothecium may be an adaptation <strong>to</strong> dispersal<br />

by arthropods.<br />

The asci are formed loosely throughout the<br />

gymnothecium. Asci are eight-spored and evanescent,<br />

releasing their ascospores passively. Ascus<br />

development is similar <strong>to</strong> that in other higher<br />

ascomycetes, with the cy<strong>to</strong>plasm being delimited<br />

by two membranes between which the ascospore<br />

wall is laid down. The inner membrane<br />

eventually becomes the ascospore plasma membrane<br />

(I<strong>to</strong> et al., 1998). The anamorphic states are<br />

usually more readily seen than the teleomorph<br />

and typically consist of rhexolytic arthrospores,<br />

although thick-walled chlamydospores are also<br />

sometimes present.<br />

Members of the Onygenales are cosmopolitan,<br />

although many individual species have a very<br />

limited distribution. Thankfully, this is true especially<br />

of many of the human pathogens. Most<br />

species, including the pathogenic ones, are soilborne<br />

and associated with keratin-containing<br />

substrates such as hair, hooves, feathers and<br />

the dung of carnivores (Hubalek, 2000). An<br />

excellent review of the order has been compiled<br />

by Currah (1985); Geiser and LoBuglio (2001) and<br />

Sugiyama et al. (2002) have discussed phylogenetic<br />

aspects.<br />

11.3.1 Onygenaceae<br />

This family contains 22 genera and 57 species<br />

and includes the most important human pathogens.<br />

The anamorphic states are arthrosporic<br />

with rhexolytic secession (e.g. Malbranchea;<br />

Fig. 11.3f), or solitary terminal spores are produced<br />

which may be unicellular (Chrysosporiumlike)<br />

or multicellular. The ascospores carry<br />

ornamentations (spines, pits or reticulations).<br />

The gymnothecia often have a few conspicuously<br />

large coiled hyphae (see Fig. 11.3a). Kwon-Chung<br />

and Bennett (1992), de Hoog et al. (2000a) and<br />

Sigler (2003) have given accounts of the most<br />

important pathogens; these are associated<br />

with the teleomorph genus Ajellomyces (Guého<br />

et al., 1997), although gymnothecia are seldom<br />

formed and the species are better known by<br />

their anamorphic names. His<strong>to</strong>plasma capsulatum,<br />

Blas<strong>to</strong>myces dermatitidis and Paracoccidioides<br />

brasiliensis are particularly closely related <strong>to</strong><br />

each other, and this grouping has been given<br />

family status by Untereiner et al. (2004), with<br />

Coccidioides immitis being less closely related and<br />

retained in the Onygenaceae. We shall consider<br />

these four pathogenic species <strong>to</strong>gether because<br />

of their medical importance. It is not permitted<br />

<strong>to</strong> work with them in standard labora<strong>to</strong>ries<br />

because they are among the handful of fungi<br />

currently listed in hazard category 3 (Kirk et al.,<br />

2001). Teleomorph genera other than Ajellomyces<br />

are Auxarthron, Amauroascus and Onygena; the<br />

latter, being the type of the family, is also briefly<br />

considered (p. 293).


ONYGENALES<br />

291<br />

Onygenaceae as human pathogens<br />

The salient features of the four important<br />

human pathogens are listed below. The points<br />

at which C. immitis differs from the other three<br />

are indicated.<br />

1. All four species are probably mainly<br />

saprotrophic in the soil, having become serious<br />

pathogens mainly because of their ability <strong>to</strong><br />

grow at 37°C, evade the human immune system,<br />

bind <strong>to</strong> human tissue, and produce proteases.<br />

Pathogenicity is probably coincidental and<br />

represents a dead end in the life cycle of these<br />

fungi because the transmission of inoculum<br />

from infected humans <strong>to</strong> the environment or<br />

<strong>to</strong> other humans is negligible (Berbee, 2001).<br />

Infection of humans occurs by inhalation of<br />

microconidia produced in the soil. These are<br />

sufficiently small <strong>to</strong> penetrate in<strong>to</strong> the alveoli of<br />

the lung. There, yeast-like stages are formed<br />

which are the agents of disease. This is in contrast<br />

<strong>to</strong> Candida albicans where hyphae rather<br />

than yeast cells represent the invasive stage.<br />

2. All species are dimorphic, with a<br />

temperature-dependent switch from hyphae<br />

(27°C) <strong>to</strong> yeast (37°C). In C. immitis, instead of<br />

producing yeast cells at 37°C, the conidium<br />

swells <strong>to</strong> produce an endospore-forming cyst or<br />

spherule. In all four species, however, the switch<br />

is relatively simple because the temperature shift<br />

is sufficient <strong>to</strong> trigger it. This differs from C.<br />

albicans in which the switch from yeast-like <strong>to</strong><br />

hyphal growth is influenced by a complexity of<br />

environmental fac<strong>to</strong>rs (see p. 277).<br />

3. Pulmonary infections may take the form<br />

of influenza-like symp<strong>to</strong>ms in the majority of<br />

immunocompetent patients but sometimes<br />

develop in<strong>to</strong> more severe tuberculosis-like<br />

illnesses. Following initial infection, yeast cells<br />

(or endospores) can be disseminated, causing<br />

systemic mycoses in other organs. In the case of<br />

mild infections, patients may make a complete<br />

recovery and may then possess lifelong immunity.<br />

This observation raises the possibility that<br />

vaccines may be developed against these pathogens<br />

(Cox & Magee, 2004; Nosanchuk, 2005) and<br />

also against Candida albicans and other fungi<br />

(Magliani et al., 2005).<br />

4. The yeast cells of P. brasiliensis,<br />

H. capsulatum and B. dermatitidis are internalized<br />

by macrophages of their human hosts, but they<br />

have a remarkable ability <strong>to</strong> survive and even<br />

reproduce inside the lytic vacuoles by raising the<br />

intravacuolar pH and withstanding the attack<br />

of the lytic enzymes and the ‘oxidative burst’<br />

created by the macrophages. Yeast cells inside<br />

macrophages represent latent inoculum which<br />

can cause relapses many years after the initial<br />

infection, especially when the host’s immune<br />

system becomes weakened by other causes. Thus,<br />

these three species have been likened, in terms of<br />

their pathology, <strong>to</strong> the bacterium Mycobacterium<br />

tuberculosis (Borges-Walmsley et al., 2002; Woods,<br />

2002).<br />

5. Even prolonged chemotherapy may not<br />

al<strong>to</strong>gether eliminate the pathogens. The drugs in<br />

common current use are similar <strong>to</strong> those applied<br />

against C. albicans (p. 278) and include amphotericin<br />

B and azole-type compounds (Harrison &<br />

Levitz, 1996). Since long treatment periods are<br />

required <strong>to</strong> control these diseases effectively,<br />

the side effects of the drugs in current use are<br />

problematic. The anti-Candida drug caspofungin<br />

(see Fig. 10.9e) also shows promise against<br />

onygenalean pathogens (Letscher-Bru &<br />

Herbrecht, 2003).<br />

6. Diseases caused by all four pathogens are<br />

much more prevalent in men than in women,<br />

often by a ratio of 10 : 1 or higher. This is due <strong>to</strong><br />

the inhibi<strong>to</strong>ry effects of oestrogen and other<br />

female steroid hormones on the conidium yeast<br />

transition (Hogan et al., 1996; Aristizabal et al.,<br />

1998).<br />

Ajellomyces capsulatus (anamorph His<strong>to</strong>plasma<br />

capsulatum)<br />

Gymnothecia of this species (Fig. 11.3a) are easily<br />

recognized, with a few conspicuous coiled<br />

appendages and very small ascospores (


292 PLECTOMYCETES<br />

His<strong>to</strong>plasma capsulatum grows as a mycelium<br />

at room temperature, but at 37°C it develops<br />

small budding yeast cells (3 5 mm) which can<br />

spread throughout the patient under favourable<br />

conditions. The normal route of infection is by<br />

inhalation, especially of the small microconidia.<br />

One common way <strong>to</strong> become infected is by<br />

cleaning buildings of bird or bat excreta, or by<br />

exploring caves in which these animals dwell;<br />

his<strong>to</strong>plasmosis is therefore also called the<br />

‘spelunker’s disease’ (Woods, 2002). Three strains<br />

of H. capsulatum are distinguished, and they are<br />

present in soil, especially when contaminated<br />

with bird or bat guano. Endemic areas are North<br />

America (var. capsulatum) or equa<strong>to</strong>rial Africa<br />

(var. duboisii). His<strong>to</strong>plasma capsulatum var. farciminosum<br />

infects horses sporadically in Africa, Asia<br />

and Eastern Europe (Weeks et al., 1985).<br />

Ajellomyces dermatitidis (anamorph<br />

Blas<strong>to</strong>myces dermatitidis)<br />

The gymnothecia of A. dermatitidis are similar <strong>to</strong><br />

those of A. capsulatus, but asexual reproduction is<br />

by means of stalked or sessile conidia which are<br />

smooth or spiny, but not tuberculate. Yeast cells<br />

are formed at 37°C, and these are much larger<br />

(10 12 mm in diameter) and have a thicker wall<br />

than those of H. capsulatum. The disease (blas<strong>to</strong>mycosis)<br />

starts as an infection of the lung<br />

which can spread systemically <strong>to</strong> other sites,<br />

especially skin and bones. As in H. capsulatum,<br />

considerable efforts are currently being made <strong>to</strong><br />

characterize the cell surface properties in<br />

B. dermatitidis (Hogan et al., 1996; Brandhorst<br />

et al., 2002). Crucial roles are probably played by<br />

the presence or absence of a-(1,3)-glucan in the<br />

cell wall, and by surface adhesins which are<br />

proteins that mediate the recognition of yeast<br />

cells by the host’s immune system.<br />

The fungus is soil-borne, especially in moist<br />

soil such as the banks of rivers. It is of North<br />

American origin, being particularly prevalent in<br />

the Mississippi and Ohio river valleys. Isolates<br />

from Africa represent a genetically different<br />

subpopulation (Guého et al., 1997).<br />

Paracoccidioides brasiliensis<br />

No teleomorph has been described as yet, but<br />

DNA sequence data predict that it will be an<br />

Ajellomyces if it is found (Guého et al., 1997).<br />

Conidia (arthroconidia) are seldom formed in<br />

culture, but the fungus is readily recognizable<br />

in the yeast state at 37°C because it produces a<br />

large, thick-walled central cell (30 mm in diameter)<br />

from which several smaller daughter cells<br />

bud off, often several at the same time. This type<br />

of budding is therefore called the ‘pilot-wheel<br />

stage’ (Fig. 11.3e). The first organ affected by<br />

paracoccidioidomycosis is the lung, but the<br />

infection may spread, causing grossly deforming<br />

lesions on mouth, nose and in gastrointestinal<br />

regions. The fungus is thought <strong>to</strong> occur in forest<br />

soils in areas with heavy rainfall in Central and<br />

South America. As with the other serious pathogens<br />

described in this section, the precise<br />

ecological niche of P. brasiliensis is still obscure,<br />

but it is possibly spread by the nine-banded<br />

armadillo, Dasypus novemcinctus (Restrepo et al.,<br />

2001).<br />

Coccidioides immitis<br />

In culture, C. immitis reproduces by Malbrancheatype<br />

arthroconidia (Fig. 11.3f). In infected tissues,<br />

arthrospores swell and give rise <strong>to</strong> large, thickwalled<br />

spherical cysts or spherules (50 100 mmin<br />

diameter) which produce endospores. Endospores<br />

are very small (3 4 mm) and are readily disseminated<br />

in the bloodstream. Each endospore can<br />

develop in<strong>to</strong> a new spherule. The disease (coccidioidomycosis)<br />

can be benign with influenza-like<br />

symp<strong>to</strong>ms, but infections of the lung may spread<br />

<strong>to</strong> other organs such as the skin, brain and bones.<br />

Coccidioidomycosis can be fatal even <strong>to</strong> immunocompetent<br />

humans (Dixon, 2001).<br />

DNA sequence comparisons indicate a relationship<br />

with the teleomorphic genus Uncinocarpus<br />

(Sigler et al., 1998). Further, detailed studies of<br />

the distribution of specific DNA sequences in the<br />

genomes of isolates from various patients have<br />

revealed that sexual recombination must occur<br />

in nature, even though no teleomorph is known<br />

(Burt et al., 1996). Coccidioides immitis is a soil-borne<br />

fungus present as arthrospores, especially in arid<br />

regions of the south-western United States and<br />

in localized places in South America. It can infect<br />

wild mammals, but outbreaks occur especially<br />

among farmers and building workers, and after<br />

dust s<strong>to</strong>rms, earthquakes and other events that


ONYGENALES<br />

293<br />

disturb the soil. One hot spot of infection is the<br />

San Joaquín Valley in Southern California, where<br />

coccidioidomycosis is known as ‘valley fever’. The<br />

Californian population of C. immitis is reproductively<br />

isolated from populations elsewhere and<br />

has recently been given the status of a separate<br />

species, C. posadasii (Fisher et al., 2002). A readable<br />

account of the his<strong>to</strong>ry of C. immitis has been given<br />

by Odds (2003), and Cox and Magee (2004) have<br />

covered general aspects of its interactions with<br />

the mammalian host.<br />

Onygena<br />

This is perhaps the most unusual yet least<br />

researched member of the Onygenaceae. Its<br />

fructification (Fig. 11.4a) is interpreted as a<br />

stalked ascostroma, i.e. an aggregate of several<br />

gymnothecia at the tip of a sterile stalk which<br />

may be 1 cm or more in length. The stalks are<br />

pho<strong>to</strong>tropic during growth. At maturity, the<br />

peridium of the stroma ruptures, thereby exposing<br />

the ascospores. There are two species,<br />

O. corvina which is associated with animal hair<br />

and bird feathers, and O. equina growing on the<br />

hooves and horns of herbivorous mammals<br />

(Currah, 1985). Both species are strongly keratinolytic,<br />

and although little work has been<br />

published on their biological features, we can<br />

assume that keratinolysis proceeds as in other<br />

Onygenales. According <strong>to</strong> Kunert (2000), the key<br />

feature is the ability <strong>to</strong> use keratin as the sole<br />

source of both carbon and nitrogen. The vast<br />

surplus of nitrogen is released in<strong>to</strong> the environment<br />

as ammonia, thereby generating an alkaline<br />

pH of 9.0 or higher. The cystein-rich keratin also<br />

contains sulphur in excess of the growth requirements<br />

of keratinolytic fungi. This is often released<br />

as sulphate, thereby buffering the pH increase<br />

caused by the release of ammonia.<br />

Like many protein-degrading fungi, Onygena<br />

produces a cadaverous smell in culture, and<br />

Currah (1985) has suggested that this might<br />

attract carrion flies if produced in nature. This<br />

would make Onygena an insect-dispersed fungus.<br />

11.3.2 Ar throdermataceae<br />

There are only two genera in this small family.<br />

Ctenomyces (Fig. 11.5), with only one species<br />

(C. serratus), has a Chrysosporium anamorph<br />

in which one-celled microconidia are formed as<br />

terminal or intercalary cells of hyphae. Species of<br />

Arthroderma also sometimes produce Chrysosporium-like<br />

microconidia (Fig. 11.6b) but are better<br />

known by their macroconidial synanamorphs<br />

Epidermophy<strong>to</strong>n, Microsporum and Trichophy<strong>to</strong>n.<br />

These are multicellular with transverse septa,<br />

and are spindle-shaped or cylindrical. They are<br />

typical and readily recognized (Fig. 11.6a). The<br />

perfect state has not been found for many<br />

of these anamorphs, but they are suspected <strong>to</strong><br />

be phylogenetically close <strong>to</strong> Arthroderma (Gräser<br />

et al., 1999; Hirai et al., 2003). Howard et al. (2003)<br />

have given a useful summary of Arthroderma and<br />

its associated anamorphs, listing 47 species. The<br />

teleomorph of the Arthrodermataceae is a typical<br />

gymnothecium with a basket of branching and<br />

anas<strong>to</strong>mosing hyphae enclosing spherical asci<br />

which release their spores passively. Characteristic<br />

appendages are often present, e.g. in<br />

Ctenomyces (Fig. 11.5), a keratinolytic species associated<br />

with feathers. The combed appendages<br />

may serve <strong>to</strong> attach the gymnothecium <strong>to</strong> bird<br />

feathers for dispersal (Currah, 1985).<br />

Members of the Arthrodermataceae are generally<br />

keratinolytic, i.e. they degrade skin and hair.<br />

Howard et al. (2003) distinguished between species<br />

primarily associated with man (anthropophilic),<br />

animals (zoophilic) or the soil (geophilic). Because<br />

of their ability <strong>to</strong> grow on the skin, hair and nails<br />

of animals, the Arthrodermataceae are collectively<br />

called derma<strong>to</strong>phytes. Diseases caused by<br />

derma<strong>to</strong>phytes are colloquially known as ‘ringworm’,<br />

whereas they are called tinea within the<br />

medical profession, with descriptive terms such<br />

as capitis, barbae, corporis and pedis added <strong>to</strong><br />

describe mycoses of the scalp, beard, general<br />

body, or feet, respectively (Howard et al., 2003).<br />

These infections are usually confined <strong>to</strong> the outer<br />

(dead) skin regions and are relatively easily<br />

controlled either by the superficial (<strong>to</strong>pical)<br />

application of creams containing a wide variety<br />

of drugs, or by oral treatment especially with<br />

triazoles or terbinafine (see p. 278; Weitzman &<br />

Summerbell, 1995; Gupta et al., 1998). Griseofulvin,<br />

produced by Penicillium griseofulvum (see<br />

p. 302), was one of the first oral and <strong>to</strong>pical<br />

drugs and is still in use <strong>to</strong>day, especially in


294 PLECTOMYCETES<br />

Fig11.5 Ctenomyces serratus. (a) Gymnothecia attached <strong>to</strong> a horse hair with which the fungus was baited. (b) Gymnothecial<br />

appendage. Note the thick walls. Micrographs taken from material kindly provided by R. Sharma.<br />

Fig11.6 Asexual states of<br />

Arthrodermaracemosum<br />

(Arthrodermataceae).<br />

(a) Microsporum-type macroconidia.<br />

The spores are multicellular and<br />

thick-walled, with granular<br />

external ornamentations.<br />

(b) Chrysosporium-like microconidial<br />

state. Redrawn from de Hoog et al.<br />

(2000a), with kind permission of<br />

Centraalbureau voor<br />

Schimmelcultures.<br />

the treatment of derma<strong>to</strong>mycoses in children,<br />

although resistance has developed among some<br />

strains of derma<strong>to</strong>phytes. Transmission of tinea<br />

infections <strong>to</strong> humans is often from pets, especially<br />

dogs, but also herbivores such as cows<br />

and horses. One species particularly common on<br />

cows is T. verrucosum (Fig. 11.7). The main<br />

transmission route for athlete’s foot and infections<br />

of <strong>to</strong>enails, both caused mainly by T.<br />

mentagrophytes and T. rubrum, is the use of<br />

communal facilities such as swimming pools<br />

and dressing rooms. Extensive listings of the<br />

various skin mycoses and the species causing<br />

them have been compiled by Weitzman and<br />

Summerbell (1995) and Howard et al. (2003).<br />

Members of the Arthrodermataceae are of<br />

his<strong>to</strong>rical interest because they were among<br />

the first fungi recognized as causing disease. In<br />

1842, Robert Remak succeeded in experimentally<br />

infecting himself with a fungus now known<br />

as T. schoenleinii (Ainsworth, 1976). Another<br />

miles<strong>to</strong>ne was laid by Raymond Sabouraud<br />

who pioneered the identification of derma<strong>to</strong>phytes<br />

on the basis of their appearance on


ONYGENALES<br />

295<br />

Fig11.7 Friesian cattle showing<br />

typical symp<strong>to</strong>ms of Trichophy<strong>to</strong>n<br />

verrucosum infections. Lesions are<br />

indicated by arrows.The circular<br />

lesion around the eye is<br />

particularly large because the<br />

animal has spread the infection<br />

by rubbing.<br />

culture media, in addition <strong>to</strong> the clinical<br />

symp<strong>to</strong>ms (see Weitzman & Summerbell, 1995).<br />

All Arthrodermataceae appear <strong>to</strong> be heterothallic.<br />

Different strains of A. simii can induce enhanced<br />

growth but not complete mating and<br />

gymnothecium formation in strains of opposite<br />

mating type of a range of other species<br />

(S<strong>to</strong>ckdale, 1968). Using this feature as a test<br />

system, it has been found for most anthropophilic<br />

species that all known isolates are of only one<br />

mating type, e.g. ( ) for T. rubrum. Since the<br />

Arthroderma state of derma<strong>to</strong>phytes is found only<br />

on substrates in contact with soil (Summerbell,<br />

2000), the lack of sexual reproduction in most<br />

anthropophilic species may be the result of<br />

adaptation <strong>to</strong> a highly patchy and specialized<br />

habitat (Howard et al., 2003).<br />

Geophilic members of the Arthrodermataceae<br />

are best isolated from soil or other substrata<br />

enriched in hair, skin or feathers (e.g. rodent<br />

burrows or birds’ nests) using horse hair or<br />

human hair as bait. Gymnothecia can be picked<br />

up with the aid of a dissection microscope,<br />

and transferred <strong>to</strong> a suitable medium such as<br />

Sabouraud agar (Sharma et al., 2002).<br />

11.3.3 Gymnoascaceae<br />

The Gymnoascaceae are a small family (10 genera,<br />

23 species; Kirk et al., 2001). A recent phylogenetic<br />

study is that by Sugiyama et al. (2002). Members<br />

of the Gymnoascaceae are isolated mainly from<br />

soil and are saprotrophic, degrading keratin and<br />

also cellulose. The genus Gymnoascus has<br />

been described by von Arx (1986). Gymnoascus<br />

reessii is a species commonly encountered on<br />

herbivore dung, producing strikingly coloured<br />

gymnothecia which are at first yellow, then red<br />

and finally brown as the ascospores mature. The<br />

reticulo peridium consists of branched, recurved,<br />

thick-walled hyphae loosely enclosing a mass of<br />

asci (Fig. 11.8). Ascocarp development can be<br />

followed readily in culture and begins from<br />

paired gametangia which arise from the same or<br />

different hyphae. The antheridium is club-shaped<br />

and the ascogonium coils around it (Fig. 11.9a).<br />

The ascogonium then becomes septate and its<br />

cells give rise <strong>to</strong> ascogenous hyphae (see Fig.<br />

11.9b), whose tips develop in<strong>to</strong> croziers. Asci<br />

develop from the penultimate cells of the croziers<br />

(Kuehn, 1956). The branched reticulo peridial<br />

hyphae arise from vegetative hyphae in the<br />

region of the gametangium (Figs. 11.9b,c). The<br />

asci do not discharge violently; the ascus wall<br />

disappears and the spores escape through the<br />

loose envelope. There is no conidial stage.<br />

11.3.4 Myxotrichaceae<br />

This family contains 4 genera (12 species).<br />

Anamorphs are Oidiodendron or Malbranchea.<br />

The genus Oidiodendron may have important


296 PLECTOMYCETES<br />

Fig11.8 Gymnoascus reessii.<br />

Ascocarp showing thick-walled<br />

branched reticulo peridial<br />

hyphae and asci.<br />

Fig11.9 Gymnoascus<br />

reessii, gymnothecium<br />

development.<br />

(a) Antheridium<br />

and ascogonium.<br />

(b) Ascogonium showing<br />

development of<br />

ascogenous hyphae.<br />

The peridial envelope<br />

is also developing.<br />

(c) Young<br />

gymnothecium showing<br />

asci at the tips of<br />

ascogenous hyphae.<br />

functions in soil ecosystems, and keys and<br />

species descriptions have been provided by<br />

Calduch et al. (2004) and Rice and Currah<br />

(2005). Myxotrichum chartarum is cellulolytic and<br />

has been known as a paper spoilage organism<br />

since the early nineteenth century. Gymnothecia<br />

form a thick chocolate-brown layer on paper<br />

or cardboard (Fig. 11.10a) and are also readily<br />

formed in culture. They are typical of the<br />

Onygenales, consisting of a reticuloperidium<br />

of anas<strong>to</strong>mosing hyphae with beautiful curved<br />

appendages (Fig. 11.10b). Other Myxotrichum spp.


EUROTIALES<br />

297<br />

have gymnothecia with simpler appendages, or<br />

these lack appendages al<strong>to</strong>gether (Currah, 1985).<br />

The taxonomic position of Myxotrichaceae is<br />

uncertain, recent phylogenetic studies suggesting<br />

an affinity with inoperculate discomycetes<br />

(Helotiales; Tsuneda & Currah, 2004). Support for<br />

this proposal comes from ecological surveys in<br />

which Oidiodendron spp. as well as members of<br />

the Helotiales have been isolated as mycorrhizal<br />

symbionts of ericaceous plants (Berch et al.,<br />

2002; Peterson et al., 2004; see also p. 442).<br />

Since both the teleomorphs and anamorphs of<br />

Myxotrichaceae are clearly referable <strong>to</strong> the<br />

Onygenales, a connection with the Helotiales<br />

would be surprising, rendering the concept of a<br />

gymnothecium with evanescent asci one of the<br />

most striking examples of convergent evolution<br />

among the fungi (Greif & Currah, 2003).<br />

11.4 Eurotiales<br />

To the practical mycology student, the order<br />

Eurotiales is among the most important groups<br />

of fungi because it contains many ubiqui<strong>to</strong>us and<br />

readily recognized species, notably in the anamorphic<br />

genera Aspergillus and Penicillium. Virtually<br />

any environmental sample soil, water, rhizosphere,<br />

air, and indoor or food contaminations<br />

will yield viable spores. One important feature of<br />

many species of Aspergillus and Penicillium is that<br />

Fig11.10 Myxotrichum chartarum. (a) Thick layer<br />

of gymnothecia on cardboard from a damp cellar.<br />

(b) Gymnothecium with dark curved appendages.<br />

Reprinted fromTribe and Weber (2002), with<br />

permission from Elsevier.


298 PLECTOMYCETES<br />

they are xerophilic, i.e. capable of growing at<br />

a water potential (a W ) at or below 0.85. Thus, they<br />

are major food spoilage organisms, growing on<br />

s<strong>to</strong>red cereals, spices, nuts, bread, dried and cured<br />

ham, pickles, jams and preserves (Lacey, 1994;<br />

Filtenborg et al., 2002). Colonization of food and<br />

feedstuff can result in its contamination by<br />

serious myco<strong>to</strong>xins (see pp. 304 306).<br />

Colonies typically spread slowly but quickly<br />

assume a greenish-blue pigmentation due <strong>to</strong><br />

abundant conidium formation. The pigmentation<br />

of mature conidia is at least partly due <strong>to</strong> melanin<br />

(see Fig. 12.46; Plate 4d). The conidial state is more<br />

commonly observed than the teleomorph, and<br />

indeed many species have lost their capacity<br />

of sexual reproduction al<strong>to</strong>gether (Geiser et al.,<br />

1996). Loss of sexual reproduction seems <strong>to</strong> have<br />

occurred independently on several occasions<br />

within the Eurotiales. When present, the ascocarps<br />

range from gymnothecial structures with<br />

loose mesh-like reticuloperidia (e.g. Talaromyces)<br />

<strong>to</strong> the hard sclerotium-like fructifications of<br />

certain species of Eupenicillium. The conidial<br />

states are generally phialidic. Microscopically,<br />

the conidiophores of Aspergillus and Penicillium<br />

can be categorized according <strong>to</strong> the arrangement<br />

of phialides on the conidiophore (Fig. 11.11). The<br />

phylogenetic value of these structures may be<br />

limited, but they are very useful for species<br />

identification. In Aspergillus (Fig. 11.11a), the conidiophore<br />

tip is swollen in<strong>to</strong> a hemispherical or<br />

club-shaped structure, the vesicle. Phialides may<br />

be formed directly at the vesicle surface in which<br />

case the conidiophore is said <strong>to</strong> be uniseriate<br />

(Figs. 11.11a, 11.16a; Raper & Fennell, 1965).<br />

Alternatively, a palisade of sterile cells (metulae)<br />

is formed by the vesicle, and the tips of the<br />

metulae give rise <strong>to</strong> the phialides (biseriate<br />

conidiophores; Fig. 11.17a). No vesicle is produced<br />

in Penicillium (Figs. 11.11b and 11.18), and instead<br />

the conidiophore tip either gives rise <strong>to</strong> phialides<br />

directly in the monoverticillate arrangement<br />

(Pitt, 1979), or it produces one series of metulae<br />

(biverticillate) or further branching layers. In<br />

terverticillate penicilli, the conidiophore tip<br />

produces one or several rami, each of which<br />

develops several metulae which in turn produce<br />

several phialides each. In quaterverticillate<br />

Fig11.11 Examples of conidiophores of three important anamorphic genera of theTrichocomaceae. (a) Aspergillus penicillioides.<br />

(b) Penicillium notatum (¼ P. chrysogenum), the original penicillin-producing strain isolated by Sir Alexander Fleming in1928.<br />

(c) Paecilomyces marquandii. All images <strong>to</strong> same scale.


EUROTIALES<br />

299<br />

penicilli, a further branch, the ramulus, is<br />

inserted between the ramus and the metula.<br />

Paecilomyces, another important anamorphic<br />

genus in the Eurotiales, has conidiophores similar<br />

<strong>to</strong> those of Penicillium, except that the phialides<br />

are more loosely arranged and have a different<br />

shape, being more elongated with a narrow<br />

drawn-out tip (Fig. 11.11c). Conidia of<br />

Paecilomyces are usually pale rather than pigmented<br />

green or blue.<br />

The order Eurotiales is thought by some <strong>to</strong><br />

consist of only a single family, Trichocomaceae,<br />

and these two terms are often used interchangeably.<br />

However, Geiser and LoBuglio (2001) also<br />

included the Monascaceae (not discussed further<br />

here) and Elaphomycetaceae (see p. 313). Many<br />

species of the Trichocomaceae are much better<br />

known by their anamorphic names, and these will<br />

continue <strong>to</strong> be used by most mycologists, especially<br />

those working in applied fields. Correlations<br />

of anamorphic and teleomorphic taxa are<br />

given in Table 11.2 (Pitt et al., 2000). These data<br />

show that a teleomorph has been found only for<br />

about 40% of all Aspergillus species described<br />

<strong>to</strong> date, and 31% of Penicillium spp. The taxonomy<br />

of the Eurotiales is still in a considerable state of<br />

confusion because it is not possible unequivocally<br />

<strong>to</strong> correlate the various anamorphs with the<br />

appearance of cleis<strong>to</strong>thecia and other features<br />

such as DNA sequence data and biochemical<br />

features (Ogawa et al., 1997; Ogawa & Sugiyama,<br />

2000).<br />

11.4.1 Aspec ts of morphogenesis<br />

in Aspergillus<br />

To recapitulate on our discussion of<br />

conidiogenesis on p. 235, a phialide, according<br />

<strong>to</strong> Kendrick (1971), is:<br />

a conidiogenous cell in which at least the first<br />

conidium initial is produced within an apical<br />

extension of the cell, but is liberated sooner or<br />

later by the rupture or dissolution of the upper<br />

wall of the parent cell. Thereafter, from a fixed<br />

conidiogenous locus, a basipetal succession of<br />

enteroblastic conidia is produced, each clad in<br />

a newly laid-down wall <strong>to</strong> which the wall of the<br />

conidiogenous cell does not contribute . . . The<br />

length of the phialide does not change during<br />

the production of a succession of conidia . . .<br />

The development of phialoconidia in Aspergillus<br />

niger is illustrated in Fig. 11.12. Young phialides<br />

are somewhat club-shaped in outline. In A. niger,<br />

A. nidulans and many other species of<br />

Aspergillus and Penicillium, the phialoconidia are<br />

uninucleate, but in some species they are multinucleate.<br />

The tip of the phialide expands <strong>to</strong> form<br />

a spherical knob which is the initial of the firstformed<br />

spore. Meanwhile, the single nucleus in<br />

the phialide divides, and a daughter nucleus<br />

passes in<strong>to</strong> the spore which begins <strong>to</strong> be<br />

separated by the formation of a septum at the<br />

phialide tip. The expansion of the first conidium<br />

leads <strong>to</strong> the rupture of the phialide wall near its<br />

tip, and the remnants of the broken phialide wall<br />

persist as a cap around the first-formed conidium.<br />

Before the rupture of the phialide wall,<br />

a layer of cell wall material is laid down<br />

(Fig. 11.12d). This layer becomes the outer wall<br />

Table 11.2. A summary of some anamorphic<br />

states found in the Trichocomaceae, and their<br />

associated teleomorphs.<br />

Anamorphic name<br />

Teleomorphic name<br />

Aspergillus (218) Chae<strong>to</strong>sar<strong>to</strong>rya (3)<br />

Emericella (33)<br />

Eurotium (24)<br />

Fennellia (3)<br />

Hemicarpenteles (2)<br />

Neosar<strong>to</strong>rya (19)<br />

Petromyces (3)<br />

Sclerocleista (2)<br />

Geosmithia (10) Talaromyces (3)<br />

Paecilomyces (48) Byssochlamys (4)<br />

Talaromyces (3)<br />

Thermoascus (4)<br />

Penicillium (249) Eupenicillium (44)<br />

Talaromyces (33)<br />

For each taxon, the numbers of species are<br />

indicated in brackets. Numerically important<br />

genera are highlighted in bold.Note that only a<br />

few ofthe 48 Paecilomycesspecies arereferable<br />

<strong>to</strong> theTrichocomaceae, many others belong <strong>to</strong><br />

Pyrenomycetes (see p. 360). Summarized from<br />

Pitt et al. (2000) and Samson (2000).


300 PLECTOMYCETES<br />

Fig11.12 Phialoconidium on<strong>to</strong>geny in Aspergillus niger (modified from Subramanian,1971). (a) Young phialide. (b) Mi<strong>to</strong>sis in phialide.<br />

(c) Conidium initial with daughter nucleus. (d) Breakage of phialide wall and formation of a new wall layer surrounding the conidial<br />

cy<strong>to</strong>plasm. (e,f) Mi<strong>to</strong>sis in the phialide and extrusion of the second phialospore. Note that the phialide has not increased in length.<br />

(g,h) Further mi<strong>to</strong>sis in the phialide and formation of the third spore. Note the formation of an inner conidial wall within the outer,<br />

and the isthmus connecting mature adjacent spores within the chain.<br />

layer of the conidium within which an inner wall<br />

develops later (Fig. 11.12h). Nuclear division<br />

continues within the phialide, and cy<strong>to</strong>plasm<br />

and a wall are laid down around a daughter<br />

nucleus <strong>to</strong> form a second conidium which is<br />

extruded from the broken tip of the phialide<br />

(Figs. 11.12e,f). The second and all subsequent<br />

conidia differ from the first in that they are not<br />

enveloped by remnants of the broken phialide<br />

wall. The cy<strong>to</strong>plasm of the second conidium is<br />

initially continuous with that of the first by<br />

means of a cylindrical isthmus. The formation of<br />

an inner wall layer by the conidia severs this<br />

cy<strong>to</strong>plasmic connection. The surviving empty<br />

isthmus is sometimes termed the connective.<br />

Each phialide can produce 100 spores or more,<br />

so that the <strong>to</strong>tal crop from one conidiophore<br />

may be more than 10 000 conidia (Adams et al.,<br />

1998). The fine structural details of conidium<br />

production in A. nidulans have been described<br />

by Mims et al. (1988).<br />

In the phialide, a transition from the polarized<br />

growth pattern of the mycelial hypha, conidiophore<br />

and phialide <strong>to</strong> yeast-like (isodiametric)<br />

growth of the nascent conidium takes place.<br />

Results obtained from experiments with the


EUROTIALES<br />

301<br />

biseriate species Aspergillus (Emericella) nidulans<br />

can be compared with those summarized earlier<br />

for the two ‘model yeasts’, Schizosaccharomyces<br />

pombe (pp. 256 259) and Saccharomyces cerevisiae<br />

(p. 270). While the cell cycle (nuclear division)<br />

is always followed by cell division (cy<strong>to</strong>kinesis)<br />

in these yeasts, this is not necessarily the case<br />

in A. nidulans. For instance, when a uninucleate<br />

conidium of A. nidulans germinates, several<br />

mi<strong>to</strong>tic divisions take place prior <strong>to</strong> and during<br />

the emergence of the germ tube before the first<br />

septum is laid down near the base of the germ<br />

tube. Subsequent septa are formed at regular<br />

intervals along the hypha, and each hyphal<br />

segment contains about 3 4 nuclei (Fiddy &<br />

Trinci, 1976). Only the nuclei in the apical cell<br />

continue <strong>to</strong> divide whereas those in intercalary<br />

compartments are arrested at the G1 stage but<br />

become reactivated later if a lateral branch is<br />

formed (Kaminskyj & Hamer, 1998). As the mycelium<br />

develops and spreads horizontally on an<br />

agar surface, conidiophores begin <strong>to</strong> grow out<br />

vertically after about 16 h in optimal conditions,<br />

with the first conidia being produced about 8 h<br />

later. Thus the time span from the germination<br />

of a conidium <strong>to</strong> the production of the new crop<br />

of conidia is only 24 h. It is not surprising that<br />

Aspergillus and Penicillium-type moulds are omnipresent<br />

in our environment!<br />

The development of the conidiophore tip has<br />

been described in detail by Mims et al. (1988) and is<br />

shown in Fig. 11.13. After vertical growth for<br />

a limited distance (about 100 mm inA. nidulans),<br />

the tip of the conidiophore swells <strong>to</strong> form the<br />

vesicle (Fig. 11.13a). In the conidiophore, cy<strong>to</strong>kinesis<br />

is suppressed whereas nuclear division<br />

continues, resulting in a multinucleate cell. In<br />

contrast, nuclear division and cy<strong>to</strong>kinesis are<br />

tightly controlled beyond the vesicle stage<br />

because a single nucleus migrates in<strong>to</strong> each of<br />

the 60 or so metulae which are formed by the<br />

vesicle of A. nidulans (Fig. 11.13b). Each metula<br />

then produces about two uninucleate phialides<br />

(Fig. 11.13c). At the phialide tip, conidia are<br />

formed (Fig. 11.13d), and here the transition<br />

from polarized <strong>to</strong> yeast-like growth takes place.<br />

The nucleus divides as the conidium enlarges, and<br />

one daughter nucleus migrates along microtubules<br />

in<strong>to</strong> the spore whereas the other remains<br />

in the phialide, ready <strong>to</strong> divide again. The signalling<br />

events which co-ordinate the interactions<br />

between the nuclear division cycle and cy<strong>to</strong>kinesis<br />

are beginning <strong>to</strong> be unravelled (Ye et al., 1999).<br />

As in the case of S. cerevisiae and S. pombe,<br />

septin-type proteins play a crucial role in organizing<br />

morphogenesis in A. nidulans and most probably<br />

other members of the Eurotiales, although<br />

detailed studies are lacking. In A. nidulans, septin<br />

rings with superimposed constricting actin rings<br />

are found at the sites of septum formation, the<br />

points of origin of metulae at the vesicle surface,<br />

the point of origin of phialides, and at<br />

the phialide tip where conidia are budded off<br />

(Momany & Hamer, 1997; Westfall & Momany,<br />

2002). All rings except for the last-mentioned are<br />

transient, disappearing as soon as the septum is<br />

complete.<br />

11.4.2 Morphogenesis in Penicillium<br />

Little work has been carried out on conidiogenesis<br />

in Penicillium, although it is reasonable <strong>to</strong> assume<br />

that the general principles will be similar <strong>to</strong> those<br />

described above for Aspergillus (Borneman et al.,<br />

2000). One interesting case is P. marneffei, which is<br />

the only species displaying a switch from hyphae<br />

<strong>to</strong> yeast cells, growing as a mycelium at 25°C and<br />

as yeast cells at 37°C. Upon transfer of a mycelial<br />

colony <strong>to</strong> 37°C, nuclear division becomes tightly<br />

coupled with cy<strong>to</strong>kinesis so that uninucleate<br />

hyphal segments are formed which fragment<br />

in<strong>to</strong> arthrospores (Garrison & Boyd, 1973;<br />

Andrianopoulos, 2002). These form yeast cells<br />

which reproduce by fission at 37°C. Interestingly,<br />

the same regula<strong>to</strong>ry mechanism is responsible<br />

for phialidic conidiogenesis at 25°C and for the<br />

switch from multinucleate hyphae <strong>to</strong> uninucleate<br />

yeast cells (Borneman et al., 2000), supporting the<br />

notion that conidiogenesis by phialides should be<br />

regarded as yeast-like growth.<br />

In Penicillium cyclopium, conidiophore formation<br />

is thought <strong>to</strong> be induced by a hormone called<br />

conidiogenone, which is permanently produced<br />

by growing colonies. When the concentration<br />

of conidiogenone exceeds a particular threshold<br />

(10 7 10 8 M), conidiogenesis, i.e. growth and<br />

differentiation of the conidiophore, is triggered<br />

(Roncal et al., 2002). For conidiogenesis <strong>to</strong> occur,<br />

hyphae must usually be exposed <strong>to</strong> air. This may


302 PLECTOMYCETES<br />

Fig11.13 Conidiophore development in Aspergillus nidulans. (a) Tip of a conidiophore which has swollen <strong>to</strong> produce a vesicle.The<br />

conidiophore is multinucleate (nuclei not visible). (b) Development of metulae. Each metula contains one nucleus. (c) Production of<br />

phialides, two from each metula.Each phialide contains one nucleus. (d) Production of uninucleate conidia. All images <strong>to</strong> same scale.<br />

be because the hormone is concentrated in the<br />

cell wall, rather than diffusing in<strong>to</strong> the medium.<br />

In liquid cultures, conidiogenesis in P. cyclopium<br />

can be triggered by the addition of Ca 2þ ions<br />

which are thought <strong>to</strong> act merely by enhancing the<br />

sensitivity of the fungus <strong>to</strong> its own hormone<br />

(Roncal et al., 2002). It is not known whether<br />

a similar hormonal system exists in Aspergillus.<br />

11.4.3 The roles of Aspergillus and Penicillium<br />

in biotechnology<br />

Species of Aspergillus and Penicillium are among<br />

the most important organisms used in biotechnology,<br />

second only <strong>to</strong> S. cerevisiae. Their applications<br />

are diverse, including the production of<br />

enzymes or primary and secondary metabolites,<br />

and direct colonization and modification of foodstuff.<br />

We will briefly consider examples of each<br />

of these applications.<br />

Food production<br />

Species of Aspergillus have been used in the Far<br />

East for food production for many centuries, and<br />

we have already mentioned the role of A. oryzae<br />

in the degradation of rice starch as a first step in<br />

saké production (see p. 276). Similar two-stage<br />

fermentation processes were developed for soy<br />

sauces, although in this case the raw material<br />

consists of a mixture of soy beans and wheat, and<br />

proteolytic as well as amylolytic enzymes are relevant.<br />

The degradation of this substrate is called<br />

the koji process, and it is one of the best examples<br />

of a solid-substrate fermentation. As such,<br />

it requires great skill <strong>to</strong> find the optimum<br />

moisture level of the soybean cake because<br />

excessive moisture will limit aeration, whereas<br />

low moisture limits growth. The koji fermentation<br />

is completed within 72 h, and it utilizes<br />

mainly A. oryzae and A. sojae. Interestingly, on the<br />

basis of morphological as well as molecular data<br />

(Geiser et al., 2000), these appear <strong>to</strong> be domesticated<br />

forms of the potent myco<strong>to</strong>xin producers<br />

A. flavus and A. parasiticus, respectively (see p. 304).<br />

The partially degraded substrate enriched in<br />

fungal extracellular enzymes is then suspended<br />

in brine, and the main fermentation (moromi) is<br />

carried out with a consortium of bacteria and<br />

yeasts. The Aspergillus enzymes continue <strong>to</strong> be<br />

active during the moromi fermentation, thus<br />

releasing a steady supply of degradation products.<br />

A readable account of soy sauce production is<br />

that by Aidoo et al. (1994). Other Far Eastern food<br />

types produced with the aid of Aspergillus spp.<br />

have been summarized by Nout and Aidoo (2002).


EUROTIALES<br />

303<br />

Cheese production involves complex consortia<br />

of bacteria and numerous different yeast species,<br />

but two Penicillium spp. are important in specialized<br />

cheeses. These are P. roqueforti for blue-veined<br />

cheeses, and P. camemberti for white mould<br />

cheeses. The subject has been summarized by<br />

Jakobsen et al. (2002). The moulds are inoculated<br />

<strong>to</strong>gether with enzymes or bacterial starters after<br />

the cooling of the pasteurized milk, and gradually<br />

colonize the maturing cheese. They contribute<br />

significantly <strong>to</strong> the texture as well as the flavours,<br />

with the characteristic blue cheese or camembert<br />

flavours being due mainly <strong>to</strong> the activity of extracellular<br />

lipases which break down short-chain<br />

fatty acids. A substantial contribution is also<br />

made by protein and peptide degradation products<br />

resulting from the activities of fungal<br />

proteinases and peptidases (Jakobsen et al., 2002).<br />

Production of enzymes<br />

The prominent role of Aspergillus and Penicillium<br />

species in food production is, of course, due <strong>to</strong><br />

their ability <strong>to</strong> produce large quantities of extracellular<br />

enzymes. This feature has also been<br />

harnessed for industrial purposes, with the<br />

majority of all commercial fungal enzymes produced<br />

by Aspergillus spp. (Oxenbøll, 1994).<br />

Proteases, amylases, lipases and pectinases are<br />

important in many industrial processes, including<br />

the manufacture of dairy, bakery, distillery<br />

and brewery products, juices and leather, and in<br />

the starch industry.<br />

Citric acid fermentation<br />

Citric acid is found in many fruits, and it is used<br />

for flavouring and pH control of food and beverages.<br />

In combination with carbonates and<br />

bicarbonates, it is also used <strong>to</strong> create the<br />

effervescent effect when medications such as<br />

vitamin preparations or aspirin are dissolved in<br />

water. Initially extracted from citrus fruits, citric<br />

acid has been produced commercially by Aspergillus<br />

niger since about 1923 and this fungus remains<br />

the world’s most important producer. The <strong>to</strong>tal<br />

current annual world production of citric acid<br />

is about 9 10 6 <strong>to</strong>ns. Curiously, although production<br />

of citric acid by A. niger is one of the most<br />

efficient biotechnological fermentations with<br />

conversion of up <strong>to</strong> 95% (by weight) of the sugar<br />

substrate, the biochemistry of it is still only<br />

poorly unders<strong>to</strong>od. The uptake of sugar (as<br />

hexose) is followed by glycolysis in the cy<strong>to</strong>sol,<br />

the tricarboxylic acid cycle in the mi<strong>to</strong>chondrion,<br />

and export of citric acid in<strong>to</strong> the cy<strong>to</strong>sol and<br />

thence in<strong>to</strong> the extracellular medium where it<br />

accumulates, creating a pH below 3. Citric acid<br />

production proceeds optimally when an excess of<br />

sugar and aeration is provided, whereas phosphate<br />

and trace elements, especially manganese,<br />

must be limiting. Excellent reviews of citric acid<br />

production have been published by Brooke (1994)<br />

and Karaffa and Kubicek (2003).<br />

Production of antibiotics<br />

The accidental discovery of penicillin by a contamination<br />

of Penicillium notatum growing on a<br />

bacterial agar culture (Fleming, 1929, 1944)<br />

followed by the re-discovery of penicillin by<br />

Florey and Chain and its development in<strong>to</strong> an<br />

antibiotic against Gram-negative bacteria has<br />

been <strong>to</strong>ld many times, and the original 1945<br />

Nobel Lectures by Fleming, Florey and Chain can<br />

be found at http://www.nobel.se. In nature,<br />

penicillin (Fig. 11.14a) is produced by P. notatum,<br />

the closely related or identical P. chrysogenum, by<br />

A. nidulans and a few other conidial fungi, whereas<br />

the chemically related cephalosporins are produced<br />

by Acremonium chrysogenum (formerly<br />

Cephalosporium chrysogenum), which probably<br />

belongs <strong>to</strong> the Pyrenomycetes (p. 348). Together,<br />

penicillin- and cephalosporin-type antibiotics<br />

take a staggering 50% share (approximately<br />

11 billion US$) of the <strong>to</strong>tal worldwide sales of<br />

antibiotics (Schmidt, 2002). Aspergillus nidulans has<br />

been useful for studies of the genetics and<br />

biosynthesis of penicillin production because<br />

it is easily manipulated by molecular biological<br />

methods. However, for commercial production<br />

P. chrysogenum has been used traditionally. The<br />

first penicillins were produced commercially<br />

by purification of the final product from<br />

static liquid cultures. Over several decades, highproducing<br />

mutants were generated, resulting<br />

in a 50 000-fold enhanced penicillin yield relative<br />

<strong>to</strong> that of the original strain, and current yields<br />

are as high as 50 g penicillin l 1 of liquid culture<br />

(Schmidt, 2002). A wide range of penicillin (and<br />

cephalosporin) derivatives has been produced,


304 PLECTOMYCETES<br />

Fig11.14 Important metabolites produced byTrichocomaceae. (a) Penicillin G, an antibiotic against Gram-positive bacteria which is<br />

synthesized from three amino acids.The cleavage point of penicillin acylases is indicated by an arrow. (b) Griseofulvin, an antifungal<br />

antibiotic which is synthesized as a heptaketide, with three methyl groups (arrows) added subsequently by methylation. (c) The<br />

polyketide afla<strong>to</strong>xin B 1<br />

. (d) Ochra<strong>to</strong>xin A.This is a pentaketide <strong>to</strong> which the amino acid phenylalanine is linked via a previously added<br />

one-carbon group (arrow). (e) The derived tetraketide patulin.<br />

partly <strong>to</strong> counter bacterial resistance and<br />

partly <strong>to</strong> broaden the range of applications or<br />

reduce allergic responses by patients. Current<br />

production seems <strong>to</strong> be mainly semi-synthetic;<br />

penicillin G is produced by P. chrysogenum,<br />

followed by the removal of the side-chain <strong>to</strong> give<br />

6-aminopenicillanic acid, which is then derivatized<br />

chemically. There are tendencies <strong>to</strong> use<br />

microbial enzymes (penicillin acylases) for the<br />

removal of the side chain (see Fig. 11.14a) and for<br />

subsequent synthetic steps (Arroyo et al., 2003;<br />

Bruggink et al., 2003). Good general reviews of<br />

the his<strong>to</strong>ry of penicillin biotechnology have<br />

been written by Rolinson (1998) and Demain and<br />

Elander (1999). The biosynthesis of penicillins<br />

and cephalosporins in fungi has been described<br />

in detail by Martin et al. (1997).<br />

Another important secondary metabolite of<br />

the Trichocomaceae is griseofulvin (Fig. 11.14b),<br />

which is used as a systemic antifungal drug,<br />

especially against derma<strong>to</strong>phytes (see p. 293).<br />

Griseofulvin was first detected in P. griseofulvum<br />

(Oxford et al., 1939) and was then re-discovered in<br />

P. janczewskii (see Brian, 1960). It is now known <strong>to</strong><br />

be produced by a wide range of Penicillium spp.<br />

as well as Aspergillus versicolor and by the Hemiascomycete<br />

Eremothecium coryli (Bérdy, 1986).<br />

Commercial production is still achieved by<br />

means of fungal fermentations.<br />

11.4.4 Myco<strong>to</strong>xins<br />

Members of the genera Aspergillus and Penicillium<br />

are no<strong>to</strong>rious for their production of secondary<br />

metabolites which are highly <strong>to</strong>xic against many<br />

different organisms and are therefore collectively<br />

called myco<strong>to</strong>xins. Since these fungi are<br />

often found as food contaminants, their myco<strong>to</strong>xins<br />

present a major health hazard and, consequently,<br />

have been thoroughly investigated.<br />

We summarize the major groups of substances<br />

here, i.e. afla<strong>to</strong>xins, ochra<strong>to</strong>xin A and patulin<br />

(Figs. 11.14c e). These are derived at least in part<br />

from the polyketide pathway in which acetylcoenzyme<br />

A or malonyl-CoA units are fused<br />

head-<strong>to</strong>-tail in a stepwise fashion. The principle


EUROTIALES<br />

305<br />

has similarities <strong>to</strong> the synthesis of fatty acids<br />

from acetyl-CoA. Synthesis proceeds in cycles,<br />

with one addition in each cycle which is followed<br />

by modification of the side chain (initially a ke<strong>to</strong><br />

group). Many other myco<strong>to</strong>xins arising from<br />

diverse biochemical pathways are produced<br />

by Aspergillus and Penicillium, and an excellent<br />

introduction <strong>to</strong> the biochemical diversity of<br />

myco<strong>to</strong>xins has been given by Moss (1994).<br />

Good reviews of ecological aspects and health<br />

implications of these myco<strong>to</strong>xins have been<br />

written by Scudamore (1994), Bhatnagar et al.<br />

(2002) and Pitt (2002). The ability <strong>to</strong> produce<br />

myco<strong>to</strong>xins such as afla<strong>to</strong>xin, ochra<strong>to</strong>xin or<br />

patulin is found in diverse groups of Aspergillus<br />

and Penicillium. Since all or most of the genes<br />

involved in the biosynthesis of a given myco<strong>to</strong>xin<br />

tend <strong>to</strong> be clustered in the genome, it is possible<br />

that sporadic horizontal gene transfer has<br />

occurred between different species, thus explaining<br />

the lack of correlation between myco<strong>to</strong>xin<br />

production and phylogenetic placement (Varga<br />

et al., 2003).<br />

Afla<strong>to</strong>xins<br />

These polyketide-type metabolites are produced<br />

by strains of Aspergillus flavus but not, apparently,<br />

by the closely related A. oryzae (Bayman &<br />

Cotty, 1993). The most common is afla<strong>to</strong>xin B 1<br />

(Fig. 11.14c), which is so named because it<br />

fluoresces blue on a thin-layer chroma<strong>to</strong>graphy<br />

plate under UV light (afla<strong>to</strong>xins G fluoresce<br />

blue green). The fluorescence of afla<strong>to</strong>xins is<br />

so strong that heavily contaminated food<br />

samples, e.g. the kernels of Brazil nuts, will<br />

fluoresce under UV light. Afla<strong>to</strong>xin B 1 is one of<br />

the most potent carcinogens known, being<br />

capable of inducing liver cancer at concentrations<br />

below 1 mgkg 1 body weight (Cotty et al.,<br />

1994). Consequently, stringent regulations<br />

concerning maximum permissible afla<strong>to</strong>xin<br />

levels are in place in many countries. However,<br />

these <strong>to</strong>xins may still present a health hazard<br />

<strong>to</strong> consumers, and also <strong>to</strong> agricultural workers<br />

because spores of A. flavus contain such high<br />

<strong>to</strong>xin levels that their inhalation may pose a risk<br />

of liver cancer (Olsen et al., 1988). Although the<br />

crop may well become contaminated on the field,<br />

A. flavus infections become visible only during the<br />

s<strong>to</strong>rage of agricultural produce. Since the fungus<br />

is xerophilic, it can colonize even dry products<br />

(Lacey, 1994). Nuts, peanuts and spices are<br />

particularly susceptible, but almost any food<br />

can be contaminated. An infamous outbreak<br />

of afla<strong>to</strong>xicosis, turkey-X disease, occurred in<br />

the UK in 1960 when about 100 000 turkeys<br />

were killed by contaminated groundnut meal.<br />

Further, cows eating contaminated feed will<br />

produce milk containing the slightly modified<br />

afla<strong>to</strong>xins M (Scudamore, 1994). Because of their<br />

ubiquity and extreme <strong>to</strong>xicity, afla<strong>to</strong>xins must<br />

be considered the most important food-borne<br />

myco<strong>to</strong>xins worldwide. The biosynthetic pathways<br />

of afla<strong>to</strong>xins are well characterized (Klich &<br />

Cleveland, 2000). The myco<strong>to</strong>xin sterigma<strong>to</strong>cystin,<br />

produced by various Aspergillus spp., is a<br />

precursor of afla<strong>to</strong>xin and is also carcinogenic,<br />

although it is comparatively rare in food and<br />

feed (Moss, 1994). The production of secondary<br />

metabolites usually occurs only when vegetative<br />

growth has ceased and when conidium formation<br />

ensues; the regula<strong>to</strong>ry mechanisms coupling<br />

conidiation with secondary metabolism are<br />

beginning <strong>to</strong> be unravelled for A. nidulans<br />

(Adams & Yu, 1998).<br />

Ochra<strong>to</strong>xin A<br />

This is a pentaketide/amino acid hybrid molecule<br />

(Fig. 11.14d) which is produced by numerous<br />

species of Aspergillus and Penicillium, especially<br />

P. verrucosum, which is common on cereals in<br />

temperate climates, and A. ochraceus and<br />

A. carbonarius, which grow on the flesh of coffee<br />

berries during drying. Coffee can, therefore, be<br />

contaminated with ochra<strong>to</strong>xin A, but mercifully<br />

much of it is destroyed during roasting (Viani,<br />

2002). Ochra<strong>to</strong>xin A consumed with contaminated<br />

cereals or meat has a long residence<br />

time (half-life 35 days) in the human body.<br />

It is highly nephro<strong>to</strong>xic and has been implicated<br />

in a degenerative human kidney disorder<br />

called ‘Balkan endemic nephropathy’; it is also<br />

strongly suspected <strong>to</strong> cause cancer of the gall<br />

bladder (S<strong>to</strong>ev, 1998; O’Brien & Dietrich, 2005).<br />

Further, ochra<strong>to</strong>xin A causes a renal degenerative<br />

disorder of farm animals, especially pigs. A<br />

chemically closely related myco<strong>to</strong>xin is citrinin.


306 PLECTOMYCETES<br />

Patulin<br />

Although patulin is a small (tetraketide-derived)<br />

molecule (Fig. 11.14e), its biosynthesis is complex,<br />

involving the formation and subsequent<br />

cleavage of an aromatic ring (Moss, 1994). Patulin<br />

is produced by several species of Aspergillus and<br />

Penicillium as well as Byssochlamys nivea (see p. 307),<br />

but the most important producer by far is<br />

P. expansum, a cause of brown rot of apples.<br />

Patulin is often detected in apple juices, sometimes<br />

at concentrations greatly exceeding safety<br />

limits set at or below 50 mgl 1 . It is, however,<br />

destroyed during alcoholic fermentation <strong>to</strong> wine<br />

or cider (Moss & Long, 2002), or by adding<br />

sulphite. It is also formed by A. clavatus in spent<br />

barley from beer brewing which is often fed <strong>to</strong><br />

cattle. Patulin may be carcinogenic; it also reacts<br />

with the sulphydryl groups of proteins, thereby<br />

inactivating enzymes (Mahfoud et al., 2002).<br />

A review of safety issues and methods for analysis<br />

and control of patulin levels in food has been<br />

written by Moake et al. (2005).<br />

11.4.5 Pathogenic species<br />

In principle, all species of Aspergillus and Penicillium<br />

and indeed many other types of fungi can<br />

cause health hazards because of the potential of<br />

their spores <strong>to</strong> act as allergens <strong>to</strong> those suffering<br />

from hay fever or asthma. Further, many species<br />

of Aspergillus and Penicillium produce myco<strong>to</strong>xins<br />

(see above). In the present section we will consider<br />

only those species which cause mycoses, i.e.<br />

infections which require chemotherapy. Good<br />

general reviews have been written by Kwon-<br />

Chung and Bennett (1992) and Summerbell<br />

(2003).<br />

Aspergillus<br />

Two species cause most of the mycotic infections<br />

associated with Aspergillus. These are A. fumigatus<br />

(69% of all reports) and A. flavus (17% of reports)<br />

(Summerbell, 2003). Both produce similar<br />

diseases. Like many other fungal pathogens of<br />

humans, these species primarily cause infections<br />

of the respira<strong>to</strong>ry tract and the lung, although<br />

wound infection can also occur occasionally. In<br />

immunocompetent patients, non-spreading<br />

‘fungus balls’ (aspergillomas) may be formed<br />

in the lung in cavities caused, for example, by<br />

previous tuberculosis. In immunocompromised<br />

patients, invasive aspergillosis may arise, i.e. the<br />

infection spreads throughout the lung and even<br />

<strong>to</strong> other organs. Aspergillosis is a major cause of<br />

death among cancer patients and is strongly on<br />

the increase among AIDS sufferers. One reason<br />

why A. fumigatus is a more frequent cause of<br />

infection than A. flavus may be that its conidia<br />

are smaller (3 mm diameter or less) and can<br />

penetrate more deeply in<strong>to</strong> the lung. They are<br />

also more buoyant in the air, and Chazalet et al.<br />

(1998) have routinely measured concentrations<br />

above 1 conidium m 3 air even in protected<br />

hospital environments. This means that every<br />

human normally inhales several hundred<br />

conidia of A. fumigatus every day. One disease<br />

caused almost solely by A. fumigatus is allergic<br />

bronchopulmonary aspergillosis, in which infections<br />

occur in patients already suffering from<br />

chronic irritation of the lung, e.g. due <strong>to</strong> asthma<br />

or cystic fibrosis. The disease can lead <strong>to</strong> fatal<br />

destruction of the lung tissue. Treatment by<br />

chemotherapy is possible, with amphotericin B<br />

and the triazole itraconazole being the major<br />

current drugs. In-depth reviews on all aspects of<br />

diseases caused by A. fumigatus have been written<br />

by Latgé (1999, 2001).<br />

Aspergillus fumigatus is a particularly thermo<strong>to</strong>lerant<br />

species with an upper growth limit at<br />

52°C (Dix & Webster, 1995), although it can survive<br />

80°C for up <strong>to</strong> 60 min (Jesenská et al., 1993).<br />

It is one of the most abundant moulds found in<br />

compost heaps and other situations in which the<br />

decay of vegetation generates heat. Workers at<br />

compost sites are therefore subjected <strong>to</strong> a massive<br />

spore inoculum, although the incidence of aspergillosis<br />

does not seem <strong>to</strong> be generally higher<br />

among them. This indicates the opportunistic<br />

nature of aspergillosis in man. In fact, the spores<br />

of thermophilic actinomycetes seem <strong>to</strong> cause<br />

most of the problems associated with ‘compost<br />

worker’s lung’ (van den Bogart et al., 1993).<br />

Penicillium<br />

In general, species of Penicillium are not as<br />

thermo<strong>to</strong>lerant as Aspergillus, with only relatively<br />

few species capable of growing at 37°C. Consequently,<br />

clinical reports of Penicillium infections


EUROTIALES<br />

307<br />

are uncommon, with the major exception of<br />

P. marneffei. As already noted, this species grows as<br />

a fission yeast at 37°C, and it causes systemic and<br />

disseminated infections in South East Asia which<br />

have increased dramatically with the spread of<br />

AIDS there. The disease symp<strong>to</strong>ms are similar<br />

<strong>to</strong> those of His<strong>to</strong>plasma capsulatum (see p. 290),<br />

including the predominance of the disease in<br />

male patients (Harrison & Levitz, 1996). Disseminated<br />

infections are most commonly found in<br />

lung, liver and skin and can be treated with<br />

amphotericin B and itraconazole (Harrison &<br />

Levitz, 1996). Penicillium marneffei can be isolated<br />

with high frequency from the internal organs of<br />

bamboo rats as well as their burrows in South East<br />

Asia. However, since contact between these<br />

rodents and humans is probably infrequent,<br />

there may be unknown sources of inoculum in<br />

the environment <strong>to</strong> which both rats and humans<br />

are exposed (Vanittanakom et al., 2006).<br />

11.4.6 Byssochlamys<br />

Byssochlamys is a small genus of soil fungi<br />

currently comprising four species (Pitt et al.,<br />

2000) which are noteworthy because of their<br />

thermo<strong>to</strong>lerance. The most <strong>to</strong>lerant structures<br />

are the ascospores which may survive heating <strong>to</strong><br />

90°C for 25 min, especially in the presence of<br />

high sucrose concentrations (Beuchat & Toledo,<br />

1977; Bayne & Michener, 1979). Byssochlamys spp.<br />

are important contaminants of canned fruits or<br />

bottled fruit juices (Tournas, 1994) because of<br />

their heat <strong>to</strong>lerance, ability <strong>to</strong> produce pec<strong>to</strong>lytic<br />

enzymes, and <strong>to</strong>lerance of conditions of low<br />

oxygen tension. Contamination can be dangerous<br />

because Byssochlamys spp. can produce several<br />

myco<strong>to</strong>xins, including patulin (Rice et al., 1977).<br />

Another habitat associated with human activity<br />

is silage in which Byssochlamys is a common<br />

contaminant (Inglis et al., 1999). In nature,<br />

Byssochlamys is ubiqui<strong>to</strong>us in the soil. In orchards,<br />

it may be splashed on<strong>to</strong> the fruit prior <strong>to</strong> or<br />

during harvesting.<br />

In culture, Byssochlamys spp. reproduce asexually<br />

by the formation of chains of hyaline conidia<br />

derived from tapering open-ended phialides<br />

(Fig. 11.15a) which have been assigned <strong>to</strong> the<br />

Paecilomyces type. Terminal, thick-walled<br />

Fig11.15 Byssochlamys nivea.<br />

(a) Phialospores and<br />

chlamydospores. (b) Coiled<br />

ascogonium surrounding an<br />

antheridium. (c) Ascogonium<br />

bearing ascogenous hyphae<br />

which in turn produce asci.<br />

Note the absence of sterile<br />

investing hyphae.


308 PLECTOMYCETES<br />

unicellular chlamydospores are also found. The<br />

asci of Byssochlamys develop best in cultures<br />

incubated around 30°C. In B. nivea, a clubshaped<br />

antheridium becomes encoiled by an<br />

ascogonium (Fig. 11.15b). Later, the coiled ascogonium<br />

develops short branches (ascogenous<br />

hyphae) which bear globose, eight-spored asci<br />

either terminally or laterally, so that eventually<br />

clusters of asci can be found. There is no sign of<br />

any sterile hyphae enclosing them (Fig. 11.15c).<br />

Like most if not all members of the Eurotiales,<br />

Byssochlamys spp. appear <strong>to</strong> be homothallic.<br />

11.4.7 Aspergillus and its teleomorphic<br />

states<br />

Keys <strong>to</strong> Aspergillus may be found in Raper and<br />

Fennell (1965), Domsch et al. (1980) and Klich<br />

(2002). There are about eight teleomorphic<br />

genera which have an Aspergillus conidial state<br />

(Table 11.2), although many Aspergillus species<br />

(about 60%) have no known sexual state. Among<br />

the purely asexual aspergilli are some of the<br />

most important species such as A. parasiticus,<br />

A. flavus, A. oryzae, A. niger and A. fumigatus. Geiser<br />

et al. (1996) proposed that the ability <strong>to</strong> reproduce<br />

sexually has been lost on many separate<br />

occasions, because many strictly mi<strong>to</strong>tic Aspergillus<br />

species have teleomorphic species as their<br />

closest relatives. The loss of the teleomorph<br />

seems <strong>to</strong> have occurred very recently on an evolutionary<br />

timescale, which is also indicated by<br />

the fact that many Aspergillus species still produce<br />

sterile structures (e.g. sclerotia) or cells<br />

(e.g. Hülle cells; see below) which are similar <strong>to</strong><br />

those found in cleis<strong>to</strong>thecia. Such defects hint<br />

at the deletion of one or more of the many genes<br />

whose products are necessary for cleis<strong>to</strong>thecium<br />

formation and meiosis. It is possible that purely<br />

asexual species are more likely <strong>to</strong> become extinct<br />

because they accumulate mutations without the<br />

possibility of meiotic recombination. Whilst the<br />

parasexual cycle can certainly be used <strong>to</strong> bring<br />

about efficient genetic recombination in the<br />

labora<strong>to</strong>ry (Bradshaw et al., 1983; see Plate 4d),<br />

it is doubtful whether it occurs sufficiently<br />

frequently in nature <strong>to</strong> present a viable alternative<br />

<strong>to</strong> meiosis (Geiser et al., 1996).<br />

It is a contentious question as <strong>to</strong> how <strong>to</strong> name<br />

purely mi<strong>to</strong>tic species. To bring current efforts at<br />

unifying anamorphic and teleomorphic taxonomy<br />

<strong>to</strong> an extreme, these Aspergillus species<br />

would have <strong>to</strong> be given the name of a teleomorph<br />

which does not exist. An additional problem is<br />

that the anamorphic features by which the<br />

genus Aspergillus is divided in<strong>to</strong> sections do not<br />

correspond <strong>to</strong> the groupings obtainable with<br />

phylogenetic analyses (Peterson, 2000b). The<br />

nomenclature and taxonomy of Aspergillus are<br />

thus in a horrible state of flux, and for the time<br />

being we have a clean conscience in continuing<br />

<strong>to</strong> use the name Aspergillus, especially for those<br />

species with no teleomorph. One desirable consequence<br />

of this approach is that it considerably<br />

reduces the degree of confusion in mycology<br />

courses. The same approach, of course, applies <strong>to</strong><br />

Penicillium. For both taxa, we will now introduce<br />

examples of the most common teleomorph<br />

forms.<br />

Eurotium<br />

Members of this genus are widely distributed in<br />

nature, especially in soil. They are responsible for<br />

the spoilage of foodstuffs, especially those with<br />

high osmotic concentrations. A typical example<br />

is E. repens (Fig. 11.16), which is common on<br />

mouldy jam. In culture, conidia are formed on<br />

agar media low in sugar content (e.g. 2% malt<br />

extract). The hyphal segment from which the<br />

conidiophore arises persists as a swollen foot cell<br />

(Fig. 11.16a). The tip of the conidiophore swells <strong>to</strong><br />

form a club-shaped vesicle bearing directly on its<br />

surface a cluster of bottle-shaped phialides which<br />

give rise <strong>to</strong> chains of green conidia in basipetal<br />

succession. On agar media with a high sugar<br />

content (e.g. 2% malt extract with 20% sucrose),<br />

yellow spherical ascocarps also develop and<br />

conidia are sparse, so that the entire Petri<br />

dish may look bright yellow instead of olive<br />

green. Aerial hyphae develop coiled ascogonia<br />

(Fig. 11.16b), and although there are reports of<br />

associated antheridia, these are not always seen.<br />

The ascogonium becomes invested by sterile<br />

hyphae which grow up from the stalk of the<br />

ascogonium. The ascogonium becomes septate,<br />

and from its segments ascogenous hyphae<br />

develop which penetrate and dissolve the


EUROTIALES<br />

309<br />

Fig11.16 Eurotium repens. (a) Conidiophore.<br />

(b)Ascogonium.(c)Ascogoniumsurroundedbysterile<br />

hyphae. (d) Cleis<strong>to</strong>thecium showing mature and<br />

immature asci.<br />

surrounding pseudoparenchyma derived from<br />

the investing hyphae. Globose asci develop from<br />

croziers at the tips of the ascogenous hyphae<br />

and, when ripe, the ascocarp consists of clusters<br />

of asci surrounded by a single-layered, yellowcoloured<br />

peridium. The peridium breaks open<br />

irregularly. The asci do not discharge violently,<br />

but the ascospores escape as the ascus wall<br />

breaks down. The ascospores are broadly lenticular,<br />

and are without obvious surface ornamentation<br />

in E. repens, although they may bear an<br />

equa<strong>to</strong>rial furrow in other species.<br />

Eurotium repens, like nearly all Eurotium spp.,<br />

is homothallic, and cultures can be transferred<br />

by means of conidia, ascospores or hyphal tips.<br />

If successive conidial transfers are made, the<br />

ability <strong>to</strong> form ascospores declines with each<br />

transfer. Ascospore production can be res<strong>to</strong>red <strong>to</strong><br />

the initial level by making one subculture from<br />

an ascospore (Mather & Jinks, 1958). This suggests<br />

that the formation of cleis<strong>to</strong>thecia and conidiophores<br />

is partially controlled by cy<strong>to</strong>plasmic<br />

determinants, and it also indicates a way<br />

in which purely anamorphic forms may have<br />

evolved from Eurotium. Little recent work seems<br />

<strong>to</strong> have been carried out on sexual reproduction<br />

in Eurotium.<br />

Emericella<br />

Emericella differs from Eurotium in a number of<br />

features. Whilst in Eurotium the ascocarp is eventually<br />

surrounded by a single-layered peridial<br />

envelope, that of Emericella is enclosed by chains<br />

of very thick-walled cells termed Hülle cells


310 PLECTOMYCETES<br />

of the nest-like arrangement of cleis<strong>to</strong>thecia surrounded<br />

by Hülle cells. This species has been used<br />

widely in genetic studies on sexual and parasexual<br />

recombination (Roper, 1966; Bos & Swart,<br />

1995) and, more recently, for experiments on<br />

morphogenetic aspects (see p. 299) as well as<br />

the biosynthetic pathways of penicillin and<br />

myco<strong>to</strong>xins (pp. 302 and 304).<br />

Fig11.17 Emericella nidulans. (a) Conidiophore. Note that the<br />

phialides are not borne directly on the vesicle. (b) Hu«lle cells,<br />

thick-walled cells surrounding the ascocarp. (c) Ascus and<br />

ascospores. Note that the ascospores bear a double flange.<br />

(a) and (c) <strong>to</strong> same scale.<br />

(Fig. 11.17b). Whereas the ascospores of Eurotium<br />

are colourless and without conspicuous ornamentations,<br />

those of Emericella are red and bear<br />

a prominent double equa<strong>to</strong>rial flange, so that<br />

the spores resemble pulley wheels (Fig. 11.17c).<br />

The conidiophores also differ in that the phialides<br />

are not borne directly on the vesicle but on<br />

a series of cylindrical cells termed metulae.<br />

The best-known species is the soil fungus<br />

Emericella (Aspergillus) nidulans, so called because<br />

11.4.8 Penicillium and its teleomorphic<br />

states<br />

Keys <strong>to</strong> Penicillium have been provided by Pitt<br />

(1979, 2000) and Ramírez (1982). The anamorphic<br />

genus Penicillium presents the same kind of taxonomic<br />

and nomenclatural problems as Aspergillus.<br />

The classical conidial apparatus is a branched<br />

conidiophore, bearing successive whorls of<br />

branches which terminate in clusters of phialides<br />

(Figs. 11.11, 11.18). Members of the subgenus<br />

Aspergilloides produce phialides directly on<br />

the conidiophore and are thus superficially similar<br />

<strong>to</strong> Aspergillus. An example is P. spinulosum<br />

(Fig. 11.18a). More commonly, the phialides are<br />

borne on a further whorl of branches, the<br />

metulae, and such species are grouped by their<br />

phialide shape and depending on whether the<br />

penicilli are symmetrical (subgenus Biverticillium)<br />

or more irregular (subgenus Furcatum). Examples<br />

are given in Figs. 11.18b,c. A third possibility<br />

is that the metulae may in turn arise from a<br />

further verticil of branches, the rami (subgenus<br />

Penicillium; e.g. P. expansum, Fig. 11.18d). In some<br />

species, e.g. P. claviforme, the individual conidiophores<br />

may be aggregated <strong>to</strong>gether in<strong>to</strong> clubshaped<br />

fructifications or coremia (Fig. 11.19).<br />

As in the case of Aspergillus, the morphology<br />

of the conidiophore unfortunately does not<br />

correlate with DNA sequencing data (Peterson,<br />

2000a), and it is possible that Penicillium will<br />

eventually be broken up in<strong>to</strong> several new genera.<br />

However, whilst of only limited taxonomic value,<br />

the conidiophore architecture will continue<br />

<strong>to</strong> be used for identification purposes. Some<br />

species of Penicillium have teleomorphs which<br />

can be assigned <strong>to</strong> Talaromyces or Eupenicillium.<br />

The different kinds of ascocarp represented by<br />

these generic names can be correlated weakly<br />

with conidiophore branching (see below). As for


EUROTIALES<br />

311<br />

Fig11.18 Conidiophores of the four anamorphic subgenera of Penicillium.(a)Penicillium spinulosum (subgenus Aspergilloides).<br />

(b) Penicillium verruculosum (subgenus Biverticillium). (c) Penicillium citrinum (subgenus Furcatum), possibly ‘one of the most common<br />

eukaryotic life forms on earth’ (Pitt,1979). (d) Penicillium expansum (subgenus Penicillium).<br />

Aspergillus, there is evidence that the ability <strong>to</strong><br />

reproduce sexually has been lost on several independent<br />

occasions in Penicillium (LoBuglio et al.,<br />

1993).<br />

Penicillium is one of the most ubiqui<strong>to</strong>us<br />

groups of fungi, occurring on all kinds of decaying<br />

materials. The conidia are universally present<br />

in air, so that Penicillium colonies are frequent<br />

contaminants of cultures. It was such a chance<br />

contaminant which led <strong>to</strong> the discovery of penicillin<br />

(see p. 302). Penicillium italicum and<br />

P. digitatum cause rotting of citrus fruits<br />

(Plate 4e) whilst P. expansum causes a brown rot<br />

of apples (see p. 304).<br />

Fig11.19<br />

plate.<br />

Penicillium claviforme producing coremia on an agar<br />

Eupenicillium<br />

Members of the genus Eupenicillium produce<br />

conidiophores which may be mono-, bi- or terverticillate.<br />

The terverticillate species (subgenus<br />

Penicillium) have been particularly well studied<br />

because many of them are relevant <strong>to</strong> man as<br />

producers of antibiotics (P. griseofulvum), in food<br />

production (e.g. P. camemberti and P. roqueforti) or


312 PLECTOMYCETES<br />

food spoilage (e.g. P. aurantiogriseum, P. digitatum,<br />

P. italicum, P. expansum), or as myco<strong>to</strong>xin producers<br />

(P. expansum). A superbly illustrated key <strong>to</strong><br />

terverticillate Penicillium spp. has been produced<br />

by Frisvad and Samson (2004), accompanied by<br />

phylogenetic analyses in which the grouping of<br />

these species in<strong>to</strong> sections has been correlated<br />

with morphological features and the production<br />

of myco<strong>to</strong>xins and other metabolites (Samson et<br />

al., 2004). An overview of secondary metabolites<br />

within Penicillium subgenus Penicillum has been<br />

compiled by Frisvad et al. (2004).<br />

In phylogenetic analyses, species of Penicillium<br />

subgenus Penicillium aggregate in a well-resolved<br />

Fig11.20 Talaromyces.(a)Talaromyces stipitatus, conidiophore. (b) Ascus and ascospores. Note the equa<strong>to</strong>rial frill. (c) Talaromyces<br />

vermiculatus, ascocarp. (d) Conidiophore. Note the long tapering phialides characteristic of the subgenus Biverticillium.<br />

(e) Ascogenous hyphae and asci. Note that some asci arise in chains.


EUROTIALES<br />

313<br />

Fig11.21 Elaphomycesgranulatus.Two<br />

ascocarps, one cut open <strong>to</strong> show<br />

contents.<br />

Fig11.22 Elaphomyces<br />

granulatus. (a) Asci showing<br />

seven and two ascospores.<br />

(b) Mature ascospore.<br />

cluster around E. crus<strong>to</strong>sum (Peterson, 2000a).<br />

Eupenicillium spp. produce cleis<strong>to</strong>thecia with<br />

very <strong>to</strong>ugh peridia. The ascospores have pulley<br />

wheel-like flanges. Penicillium spp. associated with<br />

Eupenicillium often produce thick-walled sclerotia,<br />

and it is likely that these sclerotial forms are<br />

cleis<strong>to</strong>thecial forms which have lost their ability<br />

<strong>to</strong> complete meiosis.<br />

Talaromyces<br />

The ascocarp of Talaromyces is rather different<br />

from that of Eupenicillium in having a peridium<br />

with soft cot<strong>to</strong>ny hyphae. It is thus a<br />

gymnothecium rather than a cleis<strong>to</strong>thecium.<br />

Penicillium anamorphs of Talaromyces all belong<br />

<strong>to</strong> the subgenus Biverticillium (LoBuglio et al.,<br />

1993), with long tapering phialides which are<br />

closely appressed <strong>to</strong> each other, rather than<br />

divergent (Fig. 11.20). In this they resemble<br />

Paecilomyces (Fig. 11.11c), some species of which<br />

are also associated with Talaromyces (see<br />

Table 11.2).<br />

11.4.9 Elaphomycetaceae<br />

Like so many other apparently diagnostic structures<br />

of fungi, the hypogeous (¼ subterranean)<br />

habit of the truffle has evolved independently<br />

several times. Whereas the best-known edible<br />

truffles belong <strong>to</strong> the Pezizales (see p. 423), the<br />

Elaphomycetaceae are firmly included among<br />

the Eurotiales (Geiser & LoBuglio, 2001). The<br />

genus Elaphomyces contains the most common<br />

hypogeous fungi of temperate climates, and<br />

E. granulatus (Figs. 11.21, 11.22) and E. muricatus<br />

can be collected throughout the year beneath the


314 PLECTOMYCETES<br />

litter layer under various trees, but especially<br />

beech with which they form ec<strong>to</strong>mycorrhizal<br />

associations. Elaphomyces muricatus is often parasitized<br />

by Cordyceps ophioglossoides forming yellow<br />

mycelium around the subterranean fruit bodies,<br />

and a club-shaped perithecial stroma above<br />

ground (Plate 4f). There are 5 British species<br />

(Pegler et al., 1993) and about 20 species worldwide.<br />

Elaphomyces spp., along with other truffles<br />

and epigeous fungi, form an important part<br />

of the winter diet of squirrels (Currah et al.,<br />

2000). The common name, hart’s truffle, indicates<br />

that the fruit bodies of Elaphomyces spp.<br />

are dug up and consumed by deer.<br />

The fruit bodies of Elaphomyces vary in size<br />

(about 1 4 cm in diameter) and are regarded as<br />

cleis<strong>to</strong>thecia. When cut open, a two-layered rind<br />

(peridium) can be distinguished from a central<br />

mass containing the globose asci, traversed by<br />

lighter sterile ‘veins’. The asci in E. granulatus<br />

usually contain six spores and in E. muricatus two<br />

<strong>to</strong> four. The spores are dark brown and thickwalled<br />

when mature, and the conditions necessary<br />

for their germination are not known. There<br />

are no anamorphic states.<br />

Further evidence of the remarkable plasticity<br />

of fruit body morphology in the fungi comes in<br />

the shape of an unusual member of the<br />

Elaphomycetaceae from tropical South America.<br />

In Pseudotulos<strong>to</strong>ma, a subterranean initial produces<br />

a sizeable stalk (up <strong>to</strong> 7 cm high) with<br />

a head which looks much like the gleba of<br />

a puffball but is, in fact, a cleis<strong>to</strong>thecium.<br />

At maturity, the peridial layers disintegrate,<br />

leaving the ascospores <strong>to</strong> be distributed by the<br />

wind (Miller et al., 2001).


12<br />

Hymenoascomycetes: Pyrenomycetes<br />

12.1 <strong>Introduction</strong><br />

The Pyrenomycetes are defined here according<br />

<strong>to</strong> Samuels and Blackwell (2001) as fungi which<br />

produce non-fissitunicate or occasionally pro<strong>to</strong>tunicate<br />

asci usually in flask-shaped ascomata<br />

(perithecia), less frequently in cleis<strong>to</strong>thecia. The<br />

sub-class Pyrenomycetes is one of several groups<br />

belonging <strong>to</strong> the huge and heterogeneous class<br />

Hymenoascomycetes. The characteristic feature<br />

of this class is that the asci develop in an<br />

ascohymenial way, i.e. the ascoma is formed after<br />

plasmogamy and the pairing of nuclei have<br />

occurred, and the asci therefore arise from a<br />

hymenium. This is in contrast <strong>to</strong> asci being<br />

formed singly (Archiascomycetes, Hemiascomycetes),<br />

scattered throughout the fruit body (Plec<strong>to</strong>mycetes),<br />

or formed in a locule within a preformed<br />

fruit body (Loculoascomycetes). Although<br />

the term ‘Pyrenomycetes’ is not generally unders<strong>to</strong>od<br />

in a taxonomic sense at the present,<br />

Samuels and Blackwell (2001) pointed out the<br />

monophyly of a core group of orders, including all<br />

those which we shall describe in this chapter<br />

(summarized in Table 12.1).<br />

The development of the perithecium follows<br />

several different schemes defined by Luttrell<br />

(1951), which are described in more detail for<br />

the different orders. Following fertilization and<br />

plasmogamy, the ascogonium gives rise <strong>to</strong><br />

ascogenous hyphae while the perithecial wall is<br />

formed by hyphae arising from the ascogonial<br />

stalk or elsewhere. Sterile hyphae growing up<br />

from the basal fertile region (paraphyses) and<br />

periphyses which line the inner surface of the<br />

ostiole, may be present. The development of the<br />

opening of the perithecium is typically schizogenous,<br />

i.e. it is formed by the pushing apart<br />

of tissue by the periphyses at the apex of the<br />

perithecium. This is in contrast <strong>to</strong> lysigenous<br />

development, e.g. in the pseudothecial neck of<br />

Loculoascomycetes (see p. 459). Perithecia may be<br />

formed singly or in a perithecial stroma.<br />

Ascospore discharge is by active turgor-driven<br />

liberation or occurs passively, with the ascospores<br />

oozing out of the perithecial ostiole as<br />

a tendril (cirrhus).<br />

Although the core orders belonging <strong>to</strong> the<br />

Pyrenomycetes appear <strong>to</strong> be monophyletic, the<br />

fungi considered here follow numerous different<br />

lifestyles. Many species grow saprotrophically<br />

in terrestrial habitats, while others are associated<br />

with plants, covering a wide range from<br />

mutualistic or commensalistic endophytes over<br />

biotrophic pathogens through <strong>to</strong> hemibiotrophic<br />

and necrotrophic pathogens. Animals, especially<br />

insects, are also parasitized. Numerous biologically<br />

active metabolites are produced, including<br />

alkaloids, antibiotics and phy<strong>to</strong><strong>to</strong>xins.<br />

12.2 Sordariales<br />

The order Sordariales is a substantial group of<br />

ascomycetes containing some 7 families, 115<br />

genera and over 500 species. We shall study<br />

representatives of only 2 families, the<br />

Sordariaceae (6 genera, 37 spp.) and<br />

Chae<strong>to</strong>miaceae (15 genera, 150 spp.).


316 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Table 12.1. Core groups of Pyrenomycetes as treated in this chapter. Species numbers (in brackets) are based<br />

mostly on Kirk et al. (2001).<br />

Taxon Examples ofteleomorphs Examples of anamorph<br />

Sordariales (see p. 315) Sordaria (14)<br />

Podospora and Schizothecium (80)<br />

Neurospora (12)<br />

Chrysonilia<br />

Chae<strong>to</strong>mium (80)<br />

Xylariales (see p. 332) Daldinia (13) Nodulisporium<br />

Xylaria (100)<br />

Nodulisporium<br />

Biscogniauxia (25)<br />

Nodulisporium<br />

Hypoxylon (120)<br />

Nodulisporium,Geniculosporium<br />

Kretzschmaria (5)<br />

Nodulisporium<br />

Hypocopra (30)<br />

Podosordaria (17)<br />

Lindquistia<br />

Poronia (2)<br />

Rosellinia (100)<br />

Dema<strong>to</strong>phora<br />

Hypocreales (see p. 337) Hypocrea (100) Trichoderma<br />

Hypomyces (30)<br />

Cladobotryum,Verticillium<br />

Nectria (28)<br />

Acremonium,Cylindrocarpon,<br />

Fusarium<br />

Gibberella (10)<br />

Fusarium<br />

Sphaerostilbella (4)<br />

Gliocladium<br />

Clavicipitales (see p. 348) Claviceps (36) Sphacelia<br />

Cordyceps (100)<br />

Beauveria, Metarhizium,<br />

Tolypocladium<br />

Epichloe (8)<br />

Neotyphodium<br />

Balansia (20)<br />

Ephelis<br />

Ophios<strong>to</strong>matales (see p. 364) Ophios<strong>to</strong>ma (100) Graphium, Lep<strong>to</strong>graphium,<br />

Sporothrix, Pesotum<br />

Microascales (see p. 368) Microascus (13) Cephalotrichum, Scopulariopsis,<br />

Wardomyces<br />

Cera<strong>to</strong>cystis (14)<br />

Thielaviopsis<br />

Sphaeronaemella (5)<br />

Gabarnaudia<br />

Diaporthales (see p. 373) Diaporthe (75) Phomopsis<br />

Cryphonectria (6)<br />

Endothiella<br />

Apiognomonia,Gnomoniella (18) Discula<br />

Magnaporthaceae (see p. 377) Magnaporthe (4) Pyricularia<br />

Gaeumannomyces (5)<br />

Phialophora<br />

Glomerellaceae (see p. 386) Glomerella (5) Colle<strong>to</strong>trichum<br />

The best-known genera of Sordariaceae are<br />

Sordaria, Podospora and Neurospora. Many Sordaria<br />

and Podospora spp. are coprophilous, fruiting on<br />

the dung of herbivores, but species growing on<br />

wood or in soil are also known. Neurospora occurs<br />

in nature on burnt soil and vegetation, especially<br />

in warmer countries. All three genera have<br />

been extensively used in genetic studies.


SORDARIALES<br />

317<br />

The dark-coloured perithecia usually have an<br />

ostiole lined by periphyses, but some genera are<br />

as<strong>to</strong>mous, i.e. they have fruit bodies lacking<br />

ostioles, thus forming cleis<strong>to</strong>thecia. Stromata are<br />

not produced. The asci are unitunicate and thinwalled,<br />

and the apical apparatus of the ascus is<br />

in the form of a thickened annulus or apical<br />

plate which does not stain blue with iodine. Freeended<br />

paraphyses are often present but may<br />

dissolve at ascus maturity. The ascospores are<br />

black and sometimes surrounded by a mucilaginous<br />

epispore, or they have mucilaginous<br />

appendages. The spores are mostly unicellular<br />

and germinate through a germ pore.<br />

In the Chae<strong>to</strong>miaceae, the ascomata (perithecia<br />

or sometimes cleis<strong>to</strong>thecia) are generally<br />

clothed with thick-walled ornamented hairs. The<br />

club-shaped asci are thin-walled and without<br />

apical apparatus. If a hamathecium is formed, it<br />

does not persist. The ascus wall dissolves and, in<br />

perithecial forms, the ascospores are extruded as<br />

a cirrhus. The ascospores are grey <strong>to</strong> brown in<br />

colour, mostly unicellular and with a single<br />

germ pore. Molecular data indicate a close<br />

relationship between the Chae<strong>to</strong>miaceae and<br />

Sordariaceae (Huhndorf et al., 2004).<br />

12.2.1 Sordaria (Sordariaceae)<br />

Most species of Sordaria are cellulolytic.<br />

Perithecia are common on the dung of herbivores<br />

and occasionally on other substrata such as<br />

seeds and plant remains, while a few species are<br />

reported from soil. Guarro and von Arx (1987)<br />

have given a key <strong>to</strong> 14 species, 5 of which are<br />

coprophilous. Lundqvist (1972) has described and<br />

illustrated Nordic species. Sordaria fimicola is<br />

especially common on horse dung and has been<br />

widely used in experiments on nutrition, the<br />

physiology of fruiting, spore liberation and<br />

genetics. It is homothallic, and perithecial development<br />

occurs within 10 days on a wide range of<br />

media. A longitudinal section of a perithecium<br />

(Fig. 12.1) shows a basal tuft of asci at different<br />

stages of development. The asci elongate in turn<br />

so that only one ascus can occupy the ostiole at<br />

a time. Because each spore is about 13 mm wide<br />

and the diameter of the apical apparatus of the<br />

Fig12.1 Sordaria fimicola. (a) L.S. perithecium.<br />

(b) Ascus apex. (c) Ascospore showing mucilaginous<br />

epispore. After Ingold (1971).


318 HYMENOASCOMYCETES: PYRENOMYCETES<br />

ascus is only about 4 mm, the latter acts as a<br />

sphincter, gripping the spores as they leave the<br />

ascus. Projectiles which vary in size from one <strong>to</strong><br />

eight spores may be formed. The larger the<br />

number of spores in the projectile, the greater<br />

the distance of discharge due <strong>to</strong> the fact that the<br />

surface-<strong>to</strong>-volume ratio of single spores is greater<br />

than that of multiple-spored projectiles, so that<br />

the effects of wind resistance are disproportionately<br />

high (Ingold & Hadland, 1959). Thus,<br />

single-spored projectiles have a mean discharge<br />

distance of about 1.5 cm whilst for eight-spored<br />

projectiles the distance is about 6 cm. The necks<br />

of the perithecia are pho<strong>to</strong>tropic and, as in<br />

many other coprophilous fungi, this adaptation<br />

ensures that the spores are projected away<br />

from the dung substratum. The ascospores of<br />

S. fimicola have a distinct mucilage envelope<br />

which enables them <strong>to</strong> adhere <strong>to</strong> herbage, but in<br />

S. humana the sheath is much reduced or absent.<br />

The ascospores of S. fimicola can survive for long<br />

periods. On drying, gas vacuoles (de Bary bubbles)<br />

may appear in the spore cy<strong>to</strong>plasm, but despite<br />

this the spores remain viable, and the bubbles<br />

disappear upon rehydration (Ingold, 1956;<br />

Milburn, 1970). Ascospore germination is<br />

enhanced by digestive treatment in the herbivore<br />

gut and this effect can be simulated in the<br />

labora<strong>to</strong>ry by treatment with pancreatin or<br />

sodium acetate.<br />

Perithecium development<br />

There have been numerous studies on perithecial<br />

development and structure in Sordaria, e.g. in<br />

S. fimicola (Mai, 1977), S. humana (Uecker, 1976;<br />

Read & Beckett, 1985) and S. macrospora (Hock<br />

et al., 1978). Intercalary multinucleate cells of<br />

vegetative hyphae give rise <strong>to</strong> multinucleate,<br />

spirally coiled, septate ascogonia which are not<br />

associated with antheridia. In these species,<br />

there is no trichogyne and there are no microconidia<br />

(spermatia). The ascogonium is enveloped<br />

by branched investing hyphae which<br />

originate from the ascogonial stalk or from<br />

adjacent vegetative hyphae <strong>to</strong> form a spherical<br />

pro<strong>to</strong>perithecium, i.e. an immature ascoma<br />

which does not, at this stage, show differentiation<br />

in<strong>to</strong> a neck or ostiole. The outer region of<br />

the pro<strong>to</strong>perithecium is made up of about five<br />

layers of rounded cells which have thick,<br />

pigmented (melanized) walls. Lying inside them<br />

are several layers of flattened, thinner, nonpigmented<br />

cells (see Fig. 12.1). Differentiation of<br />

the innermost cells of the pro<strong>to</strong>perithecium<br />

gives rise <strong>to</strong> the centrum consisting of elongate,<br />

free-ended, thin-walled, septate paraphyses and<br />

ascogenous cells. As the ascoma matures, it<br />

changes in outline from spherical <strong>to</strong> pearshaped<br />

with an elongate neck, thus developing<br />

in<strong>to</strong> a perithecium. Neck development is associated<br />

with meristematic activity of the cells at<br />

the base of the neck and by the appearance of<br />

short, tapering periphyses which line it. By<br />

pushing against each other the periphyses<br />

create the ostiole, the opening <strong>to</strong> the outside<br />

through which the asci will discharge their<br />

spores. Thus, the cells are pushed apart rather<br />

than lysed, and this process of ostiole development<br />

is called schizogenous. Pho<strong>to</strong>tropism of the<br />

perithecial neck is associated with differential<br />

enlargement of the periphyses.<br />

The ascogenous hyphae arise on a placentalike<br />

mound at the base of the centrum and<br />

elongate upwards. They show the usual type of<br />

crozier found in many ascomycetes with a<br />

binucleate penultimate cell (the mother cell of<br />

an ascus), a terminal cell and a stalk cell (see<br />

Fig. 8.10). Proliferation of the ascogenous hypha<br />

occurs by fusion of the recurved terminal cell<br />

with the stalk cell. Nuclear fusion in the ascus<br />

mother cell is followed by meiosis, yielding four<br />

haploid nuclei. Two further mi<strong>to</strong>tic divisions<br />

produce 16 nuclei. Cleavage of the cy<strong>to</strong>plasm<br />

accompanied by wall formation results in eight<br />

binucleate ascospores. The fine structure of<br />

ascospore development has been studied by<br />

Mainwaring (1972).<br />

Perithecium development is affected by environmental<br />

fac<strong>to</strong>rs such as temperature and<br />

nutrient supply. For S. macrospora, Hock et al.<br />

(1978) have shown that in pure culture on a<br />

defined nutrient medium, perithecium development<br />

requires a simultaneous supply of biotin<br />

and arginine. In the presence of certain other<br />

fungi such as Armillaria spp. and Mortierella spp.,<br />

S. fimicola is stimulated <strong>to</strong> increased production<br />

of perithecia and ascospores and it is likely that<br />

this effect is related <strong>to</strong> the production of


SORDARIALES<br />

319<br />

vitamins by the stimulating fungi (Watanabe,<br />

1997).<br />

Mating systems of Sordaria<br />

Although S. fimicola is homothallic, it has the<br />

ability <strong>to</strong> hybridize. Wild-type strains have black<br />

ascospores, but mutants are known with pale or<br />

colourless spores. If a wild-type strain and a<br />

white-spored mutant strain are inoculated on<br />

opposite sides of a Petri dish, hybrid perithecia<br />

develop from heterokaryotic parts of the mycelium.<br />

The asci from hybrid perithecia usually<br />

have four black and four white ascospores. Six<br />

different arrangements of the ascospores are<br />

found in such hybrid asci (Fig. 12.2). Asci with<br />

four black or four white ascospores at the tip of<br />

the ascus are those in which the gene for spore<br />

colour segregated at the first meiotic nuclear<br />

division separating the paired chromosomes. In<br />

those with two black or two white ascospores at<br />

the tip of the ascus, segregation of the gene for<br />

spore colour occurred at the second meiotic<br />

division separating the two sister chromatids of<br />

each chromosome. First-division segregation<br />

results from the absence of a cross-over between<br />

the gene for spore colour and the centromere of<br />

the chromosome, whilst second-division segregation<br />

results from a single cross-over between<br />

gene and centromere. Since the likelihood of<br />

crossing-over depends on the distance between<br />

gene and centromere, the frequency of the two<br />

kinds of segregation pattern can be used for<br />

determining the distance of the gene for spore<br />

colour relative <strong>to</strong> the centromere.<br />

A low proportion of hybrid asci shows 5 : 3,<br />

6 : 2 or (very rarely) 7 : 1 colour segregation<br />

patterns. These findings are explained in terms<br />

of gene conversion, a non-reciprocal process<br />

where one allele of a gene converts another<br />

allele at the same locus <strong>to</strong> its own type (Lamb,<br />

1996). Similar patterns have been reported<br />

in other ascomycetes, e.g. Ascobolus immersus<br />

(p. 423). A model <strong>to</strong> account for the molecular<br />

basis of gene conversion was developed with the<br />

smut fungus Ustilago maydis (see p. 652). In this<br />

so-called ‘Holliday model’ DNA strands are<br />

exchanged at ‘Holliday junctions’ between two<br />

paired DNA double helices, which can result in<br />

the generation of hybrid or heteroduplex DNA.<br />

Following separation of the two double helices<br />

from each other, non-matching DNA will be<br />

excised and the undamaged strand will be used<br />

as template <strong>to</strong> synthesize the second, damaged<br />

strand (Holliday, 1964; Lewin, 2000). This can<br />

then give rise <strong>to</strong> the phenomenon of gene<br />

conversion which, if it happens in the ascus,<br />

is manifested as deviations from the 4 : 4 gene<br />

segregation patterns.<br />

Fig12.2 Sordaria fimicola.Squash<br />

preparation from a hybrid perithecium.<br />

Most ripe asci contain four black and<br />

four white ascospores.


320 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Sordaria brevicollis and S. heterothallis are<br />

heterothallic. Both species form minute spermatia<br />

which are involved in the fertilization of<br />

mycelia of the opposite mating type. Sordaria<br />

brevicollis is a heterothallic relative of S. fimicola<br />

according <strong>to</strong> Guarro and von Arx (1987). In<br />

S. brevicollis it has been discovered that perithecium<br />

development can occur in unmated<br />

cultures of mating type A. In about 30% of such<br />

perithecia one or two asci with viable ascospores<br />

may form, with all spores being mating type A<br />

(Robertson et al., 1998).<br />

The molecular basis of homothallism has<br />

been elucidated in S. macrospora by Pöggeler<br />

et al. (1997). The molecular configuration of the<br />

genes conferring the ability <strong>to</strong> mate is similar<br />

<strong>to</strong> that in two heterothallic members of the<br />

Sordariaceae, Podospora anserina and Neurospora<br />

crassa. In Neurospora crassa (see Fig. 12.7) the<br />

haploid mating types are designated A and a. The<br />

‘alleles’ which confer mating competence in<br />

P. anserina and N. crassa consist of dissimilar<br />

DNA sequences termed idiomorphs which are<br />

present at the homologous loci in the mating<br />

partners. This term has been introduced <strong>to</strong><br />

denote sequences like those of mating types A<br />

and a, which occupy the same locus in different<br />

strains but are related neither in sequence nor<br />

(probably) by common descent (Metzenberg &<br />

Glass, 1990). In S. macrospora, the two idiomorphs<br />

are contiguous (i.e. adjoin each other), and they<br />

have been used in experiments <strong>to</strong> transform (þ)<br />

and ( ) strains of P. anserina in order <strong>to</strong> induce<br />

them <strong>to</strong> form perithecia.<br />

12.2.2 Podospora and Schizothecium<br />

(Sordariaceae)<br />

The perithecia of Podospora and Schizothecium<br />

develop on herbivore dung. In Podospora, the<br />

upper part of the perithecium wall is often<br />

ornamented by various kinds of ‘vestiture’<br />

(Lundqvist, 1972) such as short or long single<br />

hairs or long, pointed setae which may be<br />

aggregated in<strong>to</strong> a tuft <strong>to</strong> one side of the<br />

perithecial neck as seen in P. anserina and<br />

P. curvicolla. In Schizothecium, the hairs are<br />

composed of swollen cells and agglutinate<br />

<strong>to</strong>gether <strong>to</strong> form short scale-like tufts (Bell &<br />

Mahoney, 1995). The separation between these<br />

two genera has been confirmed by phylogenetic<br />

analyses (Cai et al., 2005). Taking Schizothecium<br />

and Podospora <strong>to</strong>gether, about 80 species are<br />

known (Mirza & Cain, 1969; Lundqvist, 1972).<br />

Different species show a degree of substrate<br />

specificity. For instance, perithecia of P. curvicolla,<br />

P. pleiospora and S. vesticola are especially common<br />

on the dung of lagomorphs (rabbits and hares)<br />

whilst P. curvula fruits commonly on horse dung<br />

(Lundqvist, 1972; M. Richardson, 1972, 2001). The<br />

reasons for these preferences are not known.<br />

Some species have semi-transparent perithecia<br />

within which the outline of the club-shaped asci<br />

can be seen and the sequential development<br />

and discharge of individual asci can be followed<br />

(Fig. 12.3a). The number of spores in the ascus<br />

varies from 4 <strong>to</strong> 512. Spore number has been<br />

used as a taxonomic criterion in the past,<br />

although the species concept has been widened<br />

<strong>to</strong> include forms with a range of spore numbers.<br />

For example, 8-, 16-, 32- and 64-spored forms of<br />

Podospora decipiens are recognized. The name<br />

Podospora (Gr. podos ¼ foot, spora ¼ seed) refers<br />

<strong>to</strong> the mucilaginous appendage attached <strong>to</strong> one<br />

or both ends of the black ascospore (Fig. 12.3b).<br />

This character is also found in Schizothecium. In<br />

some of the commonest species, P. curvula and<br />

S. tetrasporum, the upper spore appendages are<br />

attached <strong>to</strong> the cap of the ascus, and when the<br />

ascus explodes, the spores, roped <strong>to</strong>gether by<br />

their appendages, are propelled as a single slingshot<br />

projectile (Fig. 12.3c). As in Sordaria, it has<br />

been shown that multi-spored projectiles are<br />

discharged further than single spores (Walkey &<br />

Harvey, 1966a). The ascus wall breaks across, just<br />

beneath the cap, and in contrast <strong>to</strong> Sordaria there<br />

is usually no distinctive apical apparatus.<br />

Developmental aspects<br />

Many morphological and experimental studies<br />

have been made on P. anserina, which Lundqvist<br />

(1972) treated as a synonym of P. pauciseta. This<br />

fungus was originally described from goose dung<br />

but also fruits on the dung of sheep, horse,<br />

cattle, mice, grouse and zoo animals.<br />

Perithecium development has been described<br />

for P. anserina (Beckett & Wilson, 1968; Mai,<br />

1976) and Schizothecium spp. (Bell & Mahoney,


SORDARIALES<br />

321<br />

Fig12.3 Schizothecium tetrasporum. (a) Perithecium<br />

showing asci through the semi-transparent wall.<br />

(b) Ascus. (c) Projectile consisting of four spores<br />

attached <strong>to</strong> the ascus cap and <strong>to</strong> each other by means of<br />

mucilaginous appendages.<br />

1996). It is similar <strong>to</strong> that described above for<br />

Sordaria.<br />

The development of the ascospores of<br />

P. anserina has been studied by Beckett et al.<br />

(1968). The ascospores (four in P. anserina, but see<br />

below) are delimited by a double membrane<br />

system as in other ascomycetes (Fig. 12.4). The<br />

primary spore wall develops between the two<br />

membranes and gradually pushes them apart.<br />

The inner membrane continues <strong>to</strong> function as<br />

the plasma membrane of the spore, whilst the<br />

outer functions as the spore-investing membrane.<br />

As the primary spore wall widens,<br />

secondary wall material is laid down <strong>to</strong>wards<br />

the inside of the primary wall. These primary and<br />

secondary walls enclose the whole of the spore,<br />

including the spore head and the tail. A tertiary<br />

wall representing the pigmented layer of the<br />

spore head is laid down <strong>to</strong> the inside of the<br />

secondary layer (Figs. 12.4e g). The elongated<br />

tail of the spore is cut off from the spore head by<br />

a septum. The tertiary wall layer does not extend<br />

in<strong>to</strong> the spore tail, which therefore remains colourless.<br />

Its contents degenerate. This part of the<br />

spore persists as the primary appendage, sometimes<br />

termed the pedicel. Secondary appendages<br />

develop at the apex of the spore head<br />

and at the primary appendage. They arise<br />

by localized evaginations of the spore-investing<br />

membrane. A thinner area in the tertiary wall at


322 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.4 Podospora anserina, ascospore development (based on Beckett et al.,1968). (a) Binucleate ascospore initial enclosed by two<br />

membranes, between which the primary spore wall develops. (b d) The secondary spore wall develops within the primary wall,<br />

and the secondary appendages develop <strong>to</strong>wards each end of the spore by outpushing of the spore membrane. (e g) Development of<br />

tertiary, pigmented wall layer. A further mi<strong>to</strong>tic nuclear division occurs.The uninucleate tail of the spore is cut off from the body of<br />

the spore, and its cy<strong>to</strong>plasm degenerates, but the tail persists as the primary appendage. Note that the tertiary wall layer does not<br />

extend in<strong>to</strong> the primary appendage. At the opposite pole, a thinner area in the tertiary wall marks the position of the germ pore.<br />

the end of the spore opposite the primary<br />

appendage marks the position of the germ pore<br />

(Fig. 12.4f). The secondary appendages of many<br />

species of Podospora are very elaborate branched<br />

structures.<br />

Mating systems of Podospora<br />

Anamorphs of Podospora species, where known,<br />

consist of dark-celled phialides which produce<br />

small, sticky, unicellular, uninucleate, hyaline<br />

phialoconidia assigned <strong>to</strong> the anamorph genera<br />

Phialophora and Cladorrhinum. These do not<br />

germinate and are presumed <strong>to</strong> function as<br />

spermatia (Bell & Mahoney, 1997; Lundqvist<br />

et al., 1999).<br />

The sexual compatibility within the genus<br />

varies. Most species for which information is<br />

available are homothallic, including species with<br />

eight-spored asci such as P. decipiens, and species<br />

with more than eight spores in the ascus, e.g.<br />

P. pleiospora (Lundqvist et al., 1999). Podospora<br />

anserina and S. tetrasporum are pseudohomothallic<br />

(Esser, 1974; Raju & Perkins, 1994). Podospora<br />

anserina normally has four-spored asci, each<br />

ascospore eventually becoming quadrinucleate<br />

following two post-meiotic mi<strong>to</strong>tic nuclear divisions<br />

(see Fig. 12.6). During the final stages of<br />

development three nuclei remain in the main<br />

body of the spore and one nucleus passes in<strong>to</strong><br />

the primary spore appendage, where it


SORDARIALES<br />

323<br />

degenerates. Cultures derived from single ascospores<br />

form perithecia readily. Occasionally,<br />

however, smaller uninucleate ascospores may<br />

occur in some asci and when such spores are<br />

germinated, the resulting mycelium does not<br />

fruit. Instead, perithecia only develop when<br />

certain strains derived from uninucleate ascospore<br />

cultures are paired <strong>to</strong>gether. On each such<br />

strain, spermatia and ascogonia bearing trichogynes<br />

are formed, but these are self-incompatible;<br />

perithecia only develop if trichogynes of<br />

one strain are spermatized by spermatia of<br />

a genetically distinct strain. Thus, although the<br />

behaviour of the large ascospores suggests that<br />

P. anserina is homothallic, it is clear that the<br />

underlying mechanism controlling perithecial<br />

development is a heterothallic one of the usual<br />

bipolar type (i.e. with (þ) and ( ) strains). Most<br />

of the large ascospores (about 97%) contain<br />

nuclei of the two distinct mating types (Esser,<br />

1974).<br />

The fact that such a high proportion of the<br />

binucleate ascospores in four-spored asci contain<br />

nuclei of two distinct mating types implies some<br />

regulated process of nuclear movement and<br />

arrangement. The normal sequence of nuclear<br />

divisions occurs during ascus development,<br />

including the two nuclear divisions of meiosis<br />

and a post-meiotic mi<strong>to</strong>sis (PMM). The plane of<br />

the two meiotic nuclear divisions lies parallel <strong>to</strong><br />

the long axis of the developing ascus, but the<br />

spindles formed during PMM lie at a right angle<br />

<strong>to</strong> it (Fig. 12.5). Delimitation of the ascospores<br />

(closure) caused by invagination of the ascosporedelimiting<br />

membrane is associated with a ‘cage’<br />

of microfilaments surrounding each spore<br />

initial. At the same time a ‘rope’ of microfilaments<br />

running along the whole length of the<br />

ascus develops, and the cage of each ascospore<br />

initial becomes attached <strong>to</strong> it (Figs. 12.5a,b).<br />

Pairing between nuclei of differing mating<br />

types occurs during the cleavage of the ascospores<br />

and this appears <strong>to</strong> be mediated by astral<br />

microtubules radiating from their closely associated<br />

spindle pole bodies (SPBs) and pulling the<br />

two nuclei <strong>to</strong>gether (Thompson-Coffe & Zickler,<br />

1994). This pairing between genetically different<br />

nuclei during ascospore formation is similar<br />

<strong>to</strong> the recognition mechanism leading <strong>to</strong><br />

Fig12.5 Nuclear alignment in the pseudohomothallic<br />

four-spored fungi Podospora anserina and Neurospora<br />

tetrasperma. Fine continuous lines represent the spore cell<br />

membrane (omitted in c for clarity); dotted lines represent<br />

microfibrils; thicker continuous lines represent microtubules.<br />

Spindle pole bodies (SPBs) are shown as black bars and nuclei<br />

as open circles or ovals. (a) Following post-meiotic mi<strong>to</strong>sis<br />

(PMM) the eight nuclei are seen in pairs linked <strong>to</strong> each other<br />

by microtubules which radiate from the SPBs. A rope of<br />

actin myosin is forming along the centre line of the ascus.<br />

(b) The nuclei become rearranged and move <strong>to</strong> a staggered<br />

formation along the microfibrillar rope as microfibrillar cages<br />

form and microfibrils extend upwards <strong>to</strong>wards the SPBs.The<br />

spore membranes begin <strong>to</strong> invaginate. (c) Nuclei are re-aligned<br />

in a row, presumably by the actin myosin rope and cage<br />

assembly. Redrawn fromThompson-Coffe and Zickler (1994),<br />

with permission from Elsevier.<br />

karyogamy in the crozier prior <strong>to</strong> meiosis.<br />

Where the SPBs are not closely associated,<br />

uninucleate spores develop.<br />

It is interesting <strong>to</strong> compare nuclear behaviour<br />

within the asci of the two pseudohomothallic<br />

four-spored ascomycetes Podospora anserina and<br />

Neurospora tetrasperma (see Fig. 12.6). In N. tetrasperma,<br />

the mating type idiomorphs A and a lie<br />

close <strong>to</strong> the centromere so that A and a almost<br />

invariably go with the centromeres <strong>to</strong> opposite<br />

poles of the first meiotic division spindle (M I ),<br />

i.e. there is first-division segregation of the<br />

alleles for mating type. The second division<br />

spindles (M II ) then overlap. The third nuclear


324 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.6 Schematic diagram of ascus development in the two pseudohomothallic ascomycetes Neurosporatetrasperma and Podospora<br />

anserina (based on Raju & Perkins,1994). Neurospora tetrasperma: (a,b) First meiotic nuclear division M I<br />

.The mating type alleles A and<br />

a are closely linked <strong>to</strong> the centromere and segregate at the first division. (c) During the second meiotic nuclear division M II<br />

the<br />

spindles of the dividing nuclei lie in tandem parallel <strong>to</strong> the long axis of the ascus. (d) Interphase of M II<br />

; the nuclei are aligned in pairs.<br />

(e) Telophase of a post-meiotic mi<strong>to</strong>sis PMM I<br />

; the spindles are paired and lie transversely <strong>to</strong> the long axis of the ascus. (f) Interphase<br />

of PMM I<br />

; the four pairs of nuclei are realigned more or less regularly <strong>to</strong> one side of the ascus prior <strong>to</strong> ascospore delimitation.<br />

(g) Young ascospores are binucleate and heterokaryotic, containing both kinds of mating type allele. (h) Following a second<br />

post-meiotic mi<strong>to</strong>sis PMM II<br />

, the ascospores contain four nuclei, two of each mating type. Podospora anserina: the alleles for mating<br />

type are not tightly linked <strong>to</strong> the centromere and do not segregate at M I<br />

(b) but at M II<br />

(c). (d f) as for N. tetrasperma.(g)Following<br />

ascospore delimitation, four binucleate, heterokaryotic ascospores are formed. (h) A second post-meiotic division has occurred.<br />

Three of the four resulting nuclei, two of one mating type and one of the other, remain in the body of the ascospore whilst the<br />

fourth nucleus migrates <strong>to</strong> the basal appendage and degenerates.<br />

division (PMM) brings A and a nuclei close<br />

<strong>to</strong>gether in four pairs, and walls develop<br />

around the nuclei and cy<strong>to</strong>plasm so that the<br />

binucleate ascospores each contain one A and<br />

one a nucleus. In P. anserina, the same result is<br />

achieved by different means. The mating type<br />

idiomorphs (þ) and ( ) are sufficiently distant<br />

from the centromere for a single cross-over <strong>to</strong><br />

occur between the centromere and the locus of<br />

the idiomorphs. Second-division spindles do not<br />

overlap and, by the mechanism outlined above,<br />

nuclei of opposite mating types are brought close<br />

<strong>to</strong>gether at the PMM stage and become enclosed<br />

in walls as binucleate ascospores. A further<br />

mi<strong>to</strong>sis brings the number of nuclei in each<br />

ascospore <strong>to</strong> four, and three nuclei remain in the<br />

body of the ascospore whilst one migrates <strong>to</strong> the<br />

tail (Raju & Perkins, 1994).


SORDARIALES<br />

325<br />

Mating type fac<strong>to</strong>rs in Podospora anserina<br />

As explained above, wild-type P. anserina is<br />

pseudohomothallic, each binucleate ascospore<br />

normally containing both distinct mating type<br />

idiomorphs matþ and mat , whilst uninucleate<br />

ascospores contain either matþ or mat . The<br />

genes which control mating type specificity have<br />

been labelled FPR1 and FPR2 (fertilization plus<br />

and minus regula<strong>to</strong>rs). The molecular structure<br />

of both genes has been determined (Debuchy &<br />

Coppin, 1992; Coppin et al., 1997). The matþ locus<br />

contains 3800 + 200 base pairs (bp) with the FPR1<br />

gene within it, whilst the mat locus is larger<br />

and contains 4700 + 200 bp, enclosing FPR2 and<br />

three regula<strong>to</strong>ry genes SMR2, SMR2 and FMR1.<br />

There is a close similarity between the structure<br />

of the mating type genes in P. anserina and<br />

Neurospora crassa.<br />

Studies on incompatibility in<br />

Podospora anserina<br />

Incompatibility is usually defined as the genetic<br />

control of mating competence, but this concept<br />

extends beyond the sexual phase <strong>to</strong> the vegetative<br />

phase. Two different systems provide genetic<br />

control, namely homogenic and heterogenic<br />

incompatibility (Esser & Blaich, 1994).<br />

Homogenic incompatibility is caused by the<br />

sexual incompatibility of nuclei carrying identical<br />

idiomorphs, and it thus favours outbreeding.<br />

In contrast, in heterogenic incompatibility (also<br />

known as heterokaryon, somatic or vegetative<br />

incompatibility), the coexistence of nuclei in a<br />

common cy<strong>to</strong>plasm is inhibited by the genetic<br />

difference in one or more genes. Thus heterogenic<br />

incompatibility restricts outbreeding. It<br />

may also play a role in speciation. Another<br />

consequence is the reduced risk of transmission,<br />

following hyphal anas<strong>to</strong>mosis, of the spread of<br />

infectious cy<strong>to</strong>plasmic elements such as mycoviruses<br />

or transposons (e.g. in Cryphonectria,<br />

p. 375).<br />

Heterogenic incompatibility was discovered<br />

when attempts were made <strong>to</strong> cross strains of<br />

P. anserina of different geographic origin. When<br />

two mycelia grow <strong>to</strong>wards each other and intermingle,<br />

hyphal anas<strong>to</strong>mosis occurs. Nuclear<br />

exchange is not inhibited but is followed by an<br />

antagonistic reaction, sometimes accompanied<br />

by death of the fusing cells and by profuse<br />

branching of adjacent cells. This barrage<br />

phenomenon, observed as a white or colourless<br />

zone between two mycelia, occurs irrespective of<br />

mating type. Perithecium formation may occur<br />

in inter-racial crosses of differing mating types<br />

but the number of perithecia is much reduced. In<br />

some pairings, one or both of the reciprocal<br />

crosses between the different mating partners<br />

are unsuccessful.<br />

Nine unlinked loci are now known <strong>to</strong> be<br />

involved in the control of heterokaryon incompatibility,<br />

and these are termed het loci. A het<br />

locus can be defined as a locus in which<br />

heteroallelism cannot be <strong>to</strong>lerated in a heterokaryon.<br />

The nine het loci comprise five allelic<br />

systems (in which different alleles of the same<br />

gene provoke vegetative incompatibility) and<br />

three non-allelic systems (involving the interactions<br />

of two specific alleles from different loci).<br />

One locus (het-V) is simultaneously involved in an<br />

allelic and a non-allelic interaction (Saupe, 2000).<br />

The molecular structures of the genes at some of<br />

these loci have been characterized. The different<br />

genes encode very different products in the form<br />

of HET polypeptides. Complexes between the<br />

different HET polypeptides may function in the<br />

recognition process between self and non-self<br />

and may act as the trigger <strong>to</strong> mediate biochemical<br />

events causing vegetative incompatibility.<br />

Alternatively, HET heterocomplexes may function<br />

<strong>to</strong> poison the cell and thus may directly<br />

mediate growth inhibition and death (Glass et al.,<br />

2000).<br />

The het-s/het-S allelic system of P. anserina is of<br />

particular interest. Both alleles encode polypeptides<br />

differing only in a few amino acids,<br />

and incompatibility results if a heterodimer is<br />

formed. However, whilst the HET-S protein is<br />

immediately active, HET-s is initially translated<br />

as an inactive form, HET-s , which is present in<br />

a soluble form in the cy<strong>to</strong>plasm. Biologically<br />

active HET-s molecules may arise by a rare<br />

spontaneous rearrangement <strong>to</strong> another tertiary<br />

conformation, and HET-s molecules have the<br />

ability <strong>to</strong> convert HET-s <strong>to</strong> their own state by<br />

catalysing this conformational change. This<br />

interaction may lead <strong>to</strong> the formation of aggregates.<br />

Once initiated, the conversion of HET-s <strong>to</strong>


326 HYMENOASCOMYCETES: PYRENOMYCETES<br />

HET-s spreads like an infection throughout a<br />

mycelium at a rate of several mm h 1 . Further,<br />

transmission can occur from a HET-s containing<br />

hypha <strong>to</strong> a HET-s mycelium by anas<strong>to</strong>mosis.<br />

The ability of a protein <strong>to</strong> convert others <strong>to</strong><br />

its own state in an infectious transcriptionindependent<br />

manner, accompanied by the<br />

formation of cy<strong>to</strong>plasmic aggregates, has the<br />

hallmarks of a prion disease (Cous<strong>to</strong>u et al., 1997;<br />

Cous<strong>to</strong>u-Linares et al., 2001).<br />

Senescence in Podospora anserina<br />

Podospora anserina has been the subject of<br />

research in<strong>to</strong> senescence and has been treated<br />

as a model of the ageing phenomenon in more<br />

complex organisms (Griffiths, 1992; Bertrand,<br />

2000; Silar et al., 2001). In P. anserina senescence is<br />

defined as a diminution in the ability of cells <strong>to</strong><br />

proliferate and/or differentiate. This may or may<br />

not culminate in cell death. In pure cultures of<br />

P. anserina, senescence is marked by a progressive<br />

reduction in growth rate and loss of ability <strong>to</strong><br />

form perithecia. Eventually it proves impossible<br />

<strong>to</strong> transfer viable sub-cultures so that a given<br />

isolate has a limited lifespan. Different isolates of<br />

P. anserina have characteristic lengths of growth<br />

before growth ceases, and the mean lengths of<br />

growth can be used as a convenient indica<strong>to</strong>r of<br />

lifespan. For example, two races A and S grown<br />

on cornmeal agar in glass tubes (20 150 mm) at<br />

26°C in the dark had mean lengths of 15 and<br />

170 cm, respectively (Smith & Rubenstein, 1973).<br />

A hypothesis <strong>to</strong> explain the phenomenon of<br />

senescence in P. anserina is that, after a period of<br />

growth characteristic of a given race of the<br />

fungus, a senescence fac<strong>to</strong>r appears in a culture<br />

and is presumed <strong>to</strong> be produced, or <strong>to</strong> reproduce<br />

itself, more rapidly than other cellular components.<br />

The fac<strong>to</strong>r is transmissible through hyphal<br />

anas<strong>to</strong>mosis, i.e. fusion between a senescent<br />

hypha and a non-senescent hypha results in the<br />

non-senescent hypha acquiring the fac<strong>to</strong>r<br />

controlling senescence. Senescence is inherited<br />

maternally; it can be transmitted <strong>to</strong> 90% of the<br />

progeny of a senescent pro<strong>to</strong>perithecial strain<br />

but <strong>to</strong> none of the progeny of a spermatial<br />

parent. No nuclear mixing is involved.<br />

A large number of genes (between 600 and<br />

3000) can modulate lifespan; 50% increase it<br />

and 50% diminish it (Rossignol & Silar, 1996).<br />

Senescence is a complex process affected by<br />

environmental fac<strong>to</strong>rs and is also controlled<br />

by the interactions between nuclear and mi<strong>to</strong>chondrial<br />

DNA (Osiewacz & Kimpel, 1999;<br />

Osiewacz, 2002). The onset of senescence is<br />

marked by the appearance of dysfunctional<br />

mi<strong>to</strong>chondria and of circular plasmid-like senility<br />

DNAs derived from the mi<strong>to</strong>chondria.<br />

During the respira<strong>to</strong>ry activities of mi<strong>to</strong>chondria,<br />

reactive oxygen species (ROS) are generated<br />

as by-products and these molecules are able <strong>to</strong><br />

damage all cellular components, leading <strong>to</strong><br />

cellular dysfunctions such as the cy<strong>to</strong>chrome<br />

oxidase pathway (Osiewacz, 2002). To compensate<br />

for these dysfunctions, ROS scavengers can<br />

be produced which reduce the level of ROS.<br />

Alternative oxidative pathways may also be<br />

induced which may help in reducing the adverse<br />

effects of ROS (Dufour et al., 2000; Lorin et al.,<br />

2001). It is interesting that different mutants<br />

with defective mi<strong>to</strong>chondrial DNA associated<br />

with growth arrest can be res<strong>to</strong>red <strong>to</strong> wild-type<br />

mi<strong>to</strong>chondrial DNA by crossing, indicating<br />

an important role of sexual reproduction in<br />

this pseudohomothallic fungus (Silliker et al.,<br />

1997).<br />

12.2.3 Neurospora (Sordariaceae)<br />

There are about 12 species of Neurospora, mostly<br />

growing on soil. A key has been provided by<br />

Frederick et al. (1969). Many species grow in<br />

humid tropical and subtropical countries but<br />

others have been reported from temperate areas<br />

(Perkins & Turner, 1988; Turner et al., 2001). In<br />

nature, the most conspicuous species colonize<br />

burnt ground and charred vegetation following<br />

fire caused by volcanic eruptions or deliberate<br />

burning <strong>to</strong> clear vegetation (slash and burn) or<br />

crop residues such as sugar cane. Within a few<br />

days the burnt areas are covered by an orange or<br />

pink powdery mass of macroconidia. The<br />

commonest species here is N. intermedia (Perkins<br />

& Turner, 1988). The association with burnt<br />

ground is related <strong>to</strong> the fact that dormant<br />

ascospores in the soil are stimulated <strong>to</strong> germinate<br />

by heat. Neurospora species also grow in<br />

warm humid environments such as wood-drying


SORDARIALES<br />

327<br />

kilns and bakeries where they cause serious<br />

trouble because of their rapid growth and<br />

sporulation. For this reason N. si<strong>to</strong>phila is sometimes<br />

called the red bread mould. Strains of<br />

N. intermedia are used in the preparation of the<br />

fermented food ‘omchom’ (ontjom) in which<br />

conidia are used <strong>to</strong> inoculate soybean or<br />

peanut solids from which oil and protein have<br />

been extracted by pressing.<br />

Neurospora as a genetic <strong>to</strong>ol<br />

Neurospora has been widely used in genetic and<br />

biochemical studies (Perkins, 1992; Davis, 1995).<br />

The development of auxotrophic mutants<br />

deficient in successive steps of arginine<br />

biosynthesis led <strong>to</strong> the proposition of the ‘onegene<br />

one-enzyme’ hypothesis by Beadle and<br />

Tatum (1941) who were awarded the Nobel<br />

Prize in 1958. There is a very extensive literature<br />

relating <strong>to</strong> the genus. The loci of over 1000 genes<br />

have been mapped (Perkins et al., 2000), and the<br />

complete genome of N. crassa has now been<br />

sequenced. It is haploid and has seven chromosomes.<br />

The best-known species are N. crassa and<br />

N. si<strong>to</strong>phila, both of which are eight-spored and<br />

heterothallic. Neurospora tetrasperma is fourspored<br />

and pseudohomothallic. Some other<br />

species, e.g. N. africana, N. dodgei and N. terricola,<br />

are homothallic (Coppin et al., 1997). The homothallic<br />

species do not form conidia.<br />

The reasons why Neurospora has proven so<br />

useful as a <strong>to</strong>ol in biochemical and genetic<br />

research are: (1) that it is haploid; (2) that wildtype<br />

strains have simple nutritional requirements,<br />

namely a carbon source, simple mineral<br />

salts and one vitamin, biotin; (3) that mutations<br />

can be induced readily by the use of chemical<br />

mutagens or UV irradiation of conidia; (4) that<br />

growth and sexual reproduction is rapid; and<br />

(5) that tetrad analysis by dissecting asci is<br />

straightforward. By the use of marked strains it<br />

is now not even necessary <strong>to</strong> dissect asci because<br />

tetrad analysis can be performed on octets of<br />

projected ascospores. Another advantage of<br />

Neurospora is that cultures can be preserved for<br />

long periods in suspended animation as spores<br />

s<strong>to</strong>red over silica gel, or following lyophilization<br />

or freezing.<br />

The life cycle of Neurospora<br />

The life cycle of N. crassa is illustrated diagrammatically<br />

in Fig. 12.7. The name Neurospora is<br />

derived from the characteristically ribbed ascospores.<br />

The ascospore walls bear dark, raised,<br />

thicker ribs separated by thinner, paler,<br />

branched or unbranched inter-costal veins<br />

(Figs. 1.19 and 12.8). The spores are multinucleate<br />

and have reserves of lipids and the carbohydrate<br />

trehalose. The dark ascospores of N. crassa are<br />

viable for many years and do not germinate<br />

readily unless treated chemically (e.g. by furfural)<br />

or by heat shock (e.g. 60°C for 20 40 min).<br />

In contrast, the conidia are killed by such heat<br />

treatment. Following treatment, the ascospores<br />

germinate through a germ pore at either or both<br />

ends, forming at first a globose, inflated vesicle<br />

and then a coarse, incompletely septate, rapidly<br />

growing mycelium, each segment of which is<br />

multinucleate. Cy<strong>to</strong>plasm and organelles including<br />

nuclei pass freely through the septa.<br />

Within 24 h, the mycelium can begin asexual<br />

reproduction. The cues inducing conidial development<br />

include light, desiccation, and nutrient<br />

deprivation. Upright branches develop which,<br />

instead of continuing <strong>to</strong> grow by hyphal tip<br />

elongation, undergo repeated apical budding.<br />

The resulting cells are separated from each other<br />

by incomplete septa with a wide central pore.<br />

These cells have been termed proconidia<br />

(Springer & Yanofsky, 1989). The proconidia<br />

continue <strong>to</strong> bud apically, forming multinucleate<br />

macroconidia (blas<strong>to</strong>conidia) separated from<br />

each other by septa with narrower pores. The<br />

septa thicken and develop a centripetal furrow<br />

which widens, leaving a central strand of<br />

material, the connective, by which the spores<br />

remain attached <strong>to</strong> each other. Further conidia<br />

develop by budding of the terminal conidium<br />

on a chain and, when the terminal conidium<br />

gives rise <strong>to</strong> two buds, the chain branches<br />

(Figs. 12.8d,e; Hashmi et al., 1972). Conidia of<br />

this type belong <strong>to</strong> the form genus Chrysonilia<br />

(formerly Monilia). The individual segments of the<br />

spore chain break apart and are readily dispersed<br />

by wind. The surface of macroconidia in wildtype<br />

strains is made up of hydrophobin rodlets<br />

which impregnate it, but the conidia of the<br />

easily wettable mutant eas are devoid of rodlets


328 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.7 The life cycle of Neurospora crassa, diagrammatic and not <strong>to</strong> scale (based on Fincham & Day,1971).The fungus is<br />

heterothallic and a mature ascus (left of diagram) contains four haploidmultinucleate ascospores of mating type a and four of mating<br />

type A. Ascospores germinate <strong>to</strong> form a branched, incompletely septate mycelium with multinucleate segments. Multinucleate<br />

macroconidia and uninucleate microconidia develop, and both types of conidium can germinate <strong>to</strong> form a new mycelium.<br />

Pro<strong>to</strong>perithecia, consisting of a coiled ascogonium and a trichogyne, and surrounded by sterile hyphae, also develop on the haploid<br />

mycelium (right of diagram).When microconidia or macroconidia of opposite mating type are transferred <strong>to</strong> a trichogyne,<br />

plasmogamy (P) occurs and the pro<strong>to</strong>perithecium develops in<strong>to</strong> a perithecium. Ascogenous hyphae containing paired A and a nuclei<br />

grow out from the ascogonium. At the tip of the ascogenous hypha a crozier develops (bot<strong>to</strong>m of diagram) and, within the<br />

penultimate cell, karyogamy (K) occurs between nuclei of the two mating types.Within this diploid cell, which is the ascus initial,<br />

meiosis (M) takes place, followed by mi<strong>to</strong>ses (not shown). Ascospores are cleaved around the nuclei (not shown) and the mature<br />

asci discharge their ascospores through the neck of the perithecium.


SORDARIALES<br />

329<br />

Fig12.8 Neurospora crassa. (a) Ascus.<br />

(b) Ascospores showing ribbed surface.<br />

(c) Pro<strong>to</strong>perithecium showing projecting<br />

trichogyne. (d) Macroconidia from one-day-old<br />

culture. (e) Enlarged view of developing<br />

macroconidia. (f) Microconidia forming sticky<br />

clusters. (g) Enlarged view showing origin of<br />

microconidia. (a d,f) <strong>to</strong> same scale;<br />

(e,g) <strong>to</strong> same scale.<br />

and cohere in sticky masses (Beever et al., 1978).<br />

The pink colour of the conidia is due <strong>to</strong> the<br />

presence of a carotenoid pigment, neurosporoxanthin,<br />

which is stimulated <strong>to</strong> develop by<br />

light. The spores are formed in vast numbers.<br />

In labora<strong>to</strong>ries they can cause serious contamination<br />

of other cultures, partly because of the<br />

rapid growth of the mycelium (up <strong>to</strong> 5 mm h 1 ),<br />

and partly because macroconidia can develop<br />

in profusion beyond the rims of closed Petri<br />

dishes. In humid environments, the mycelium<br />

may grow through the cot<strong>to</strong>n wool plugs of testtube<br />

cultures and sporulate on the plugs. Shaw<br />

(1993) has reported that macroconidia formed on<br />

steamed pine logs in Queensland, Australia, are<br />

collected by honeybees in their pollen baskets.<br />

Cultures derived from a single ascospore also<br />

develop two other types of reproductive structure.<br />

In contrast <strong>to</strong> the large, dry, wind-dispersed<br />

macroconidia which are formed within 1 2 days,<br />

clumps of smaller, oval, uninucleate, sticky<br />

microconidia develop after about 12 15 days<br />

(Figs. 12.8f,g; Maheshwari, 1999). The conidiogenous<br />

cells from which the microconidia develop<br />

have been interpreted as reduced phialides. The<br />

microconidia are capable of slow and variable<br />

germination but function primarily as spermatia.<br />

Coiled ascogonia, terminated by long tapering<br />

trichogynes and surrounded at the base by<br />

hyphae, also develop (Fig. 12.8c). Such structures<br />

are termed pro<strong>to</strong>perithecia or bulbils and each<br />

often forms several trichogynes. In N. crassa and<br />

N. si<strong>to</strong>phila, no further development occurs in<br />

single ascospore cultures, i.e. each strain is selfincompatible.<br />

Incompatibility is controlled by a<br />

pair of mating type idiomorphs, A and a, and<br />

if two compatible strains are grown <strong>to</strong>gether in<br />

a Petri dish for a few days, microconidia of one


330 HYMENOASCOMYCETES: PYRENOMYCETES<br />

strain can be transferred <strong>to</strong> trichogynes of the<br />

opposite strain by flooding with sterile water. In<br />

nature it is possible that mites or insects are<br />

involved in transfer. There is evidence that a<br />

microconidium produces a pheromone which<br />

induces directional growth (positive chemotropism)<br />

of a trichogyne of opposite mating type<br />

<strong>to</strong>wards it before plasmogamy occurs (Bistis,<br />

1983). The transfer of macroconidia or hyphae<br />

of the opposite strain <strong>to</strong> a trichogyne can also<br />

effect fertilization. Fusion between the trichogyne<br />

and the fertilizing cell is followed by<br />

the migration of one or more nuclei from the<br />

fertilizing cell down the trichogyne in<strong>to</strong> the<br />

ascogonium. The development of ripe perithecia<br />

occurs within 7 10 days and follows the typical<br />

general ascomycete pattern. This has been<br />

described by Nelson and Backus (1968) in two<br />

homothallic species. Much is now known of the<br />

genetic control of sexual development in<br />

Neurospora and several genes have been identified<br />

which control steps in the process. The cy<strong>to</strong>logical<br />

details of ascus development have also<br />

been worked out. After the eight-nucleate<br />

stage in the developing ascus of N. crassa, several<br />

mi<strong>to</strong>ses ensue so that a fully developed<br />

ascospore may contain as many as 32 nuclei<br />

(Raju, 1992a).<br />

Mating type genes in Neurospora<br />

In N. crassa, heterokaryons are not normally<br />

formed between mycelia of opposite mating<br />

types, and this implies that plasmogamy usually<br />

occurs only between a trichogyne of one strain<br />

and a fertilizing agent (e.g. a microconidium,<br />

macroconidium or hypha) of the opposite strain.<br />

This condition is termed restricted heterokaryosis.<br />

Even within one mating type, the ability<br />

<strong>to</strong> form heterokaryons is under genetic control,<br />

i.e. there is heterokaryon incompatibility as in<br />

Podospora anserina and heterokaryons generally<br />

develop only if the het genes which control<br />

compatibility are homoallelic (see p. 325).<br />

Several het genes have been identified (Mylyk,<br />

1976; Perkins, 1992), and Micali and Smith (2003)<br />

have provided evidence of a yet more complex<br />

regulation of heterokaryon incompatibility in<br />

the shape of suppressor genes which modify the<br />

effect of het and mating type genes. When the<br />

mycelia of unlike genotype anas<strong>to</strong>mose, cy<strong>to</strong>plasmic<br />

incompatibility results in vacuolation<br />

and disorganization of cell contents in the<br />

region of the anas<strong>to</strong>mosis. Similar cy<strong>to</strong>plasmic<br />

reactions are visible when anas<strong>to</strong>mosis occurs<br />

between the hyphae of wild-type strains differing<br />

in mating types. In contrast <strong>to</strong> N. crassa, heterokaryons<br />

are readily formed between different<br />

mating type strains of N. tetrasperma, which thus<br />

exhibits unrestricted heterokaryosis.<br />

The molecular structure of the A and a<br />

idiomorphs has been elucidated in N. crassa.<br />

They are strikingly dissimilar (Glass et al., 1988).<br />

The A idiomorph is composed of a region of<br />

5301 bp bearing little similarity <strong>to</strong> the a idiomorph<br />

comprising 3235 bp (Glass et al., 1990;<br />

Staben & Yanofsky, 1990). The A idiomorph of<br />

N. crassa is also involved in heterokaryon incompatibility.<br />

There are similarities in the structure<br />

and functions of the mating type idiomorphs<br />

between N. crassa, Podospora anserina and the<br />

yeasts Saccharomyces cerevisiae and<br />

Schizosaccharomyces pombe, but there are also<br />

differences (p. 266; Glass & Lorimer, 1991). The<br />

idiomorphs of N. crassa are larger than those of<br />

the yeasts. Mating type idiomorphs are present<br />

in several homothallic species of Neurospora<br />

which hybridize with the A DNA probe of N.<br />

crassa (Coppin et al., 1997). This and other lines of<br />

evidence suggest that the heterothallic condition<br />

was primitive and that the homothallic condition<br />

has probably arisen several times in the<br />

course of evolution and may have a selective<br />

advantage (Metzenberg & Glass, 1990).<br />

The pseudohomothallic condition, represented<br />

by N. tetrasperma, is of interest. This<br />

species is functionally homothallic in that the<br />

mycelium from a single heterokaryotic ascospore<br />

containing nuclei of two distinct mating types<br />

can develop perithecia directly. However, homokaryotic<br />

mycelia can develop from uninucleate<br />

ascospores and also from about 20% of macroconidia.<br />

Such mycelia are capable of outcrossing.<br />

Raju (1992b) has summarized,<br />

Thus N. tetrasperma appears <strong>to</strong> have the best of<br />

both worlds. On one hand the single-mating-type<br />

homokaryotic cultures offer N. tetrasperma the<br />

advantages of outbreeding. On the other hand the


SORDARIALES<br />

331<br />

heterokaryotic (A þ a) cultures have the potential<br />

advantage of hybrid vigour. They can also enter the<br />

sexual cycle and produce ascospores without waiting<br />

for a sexual partner. This is very useful in situations<br />

where rapid completion of the life cycle is<br />

advantageous.<br />

Neurospora and the biological clock<br />

Many organisms show a daily (circadian) rhythm<br />

in their activities, entrained by a combination<br />

of light and temperature (Dunlap, 1999). When<br />

transferred <strong>to</strong> a uniform environment, they may<br />

continue <strong>to</strong> display the same rhythm, suggesting<br />

that it is controlled by some internal clock.<br />

Important research helping <strong>to</strong> interpret the<br />

molecular basis of circadian rhythmicity in<br />

fungi as well as other eukaryotes is being<br />

performed on N. crassa (Davis, 1995; Liu, 2003;<br />

Dunlap & Loros, 2004). A band mutant (bd) was<br />

discovered which, when grown in culture in long<br />

tubes, formed alternating bands of macroconidia<br />

interspersed by non-sporulating bands. In continuous<br />

darkness, bands continued <strong>to</strong> form, with a<br />

periodicity of 21.5 h, little affected by temperature.<br />

From the bd mutant, further mutants were<br />

developed by mutagenesis in which the frequency<br />

of the circadian rhythm was affected. The frq<br />

locus was identified as a key control element<br />

of the frequency of sporulation, with partial<br />

loss-of-function mutations capable of shortening<br />

the frequency <strong>to</strong> as little as 16 h or extending<br />

it <strong>to</strong> 29 h. The effect of temperature on the<br />

circadian rhythm seems <strong>to</strong> be controlled by the<br />

temperature-dependent alternative splicing of<br />

the frq mRNA, giving either of two major FRQ<br />

proteins (Colot et al., 2005). Several other loci are<br />

also involved in the integration of the rhythm<br />

with temperature and light. The white collar wc-1<br />

and wc-2 gene products are especially important,<br />

forming a heterodimer (WCC ¼ white collar<br />

complex) which stimulates frq transcription. The<br />

FRQ dimer, in turn, inhibits existing WCC and<br />

lowers the continued expression of WC-1, thus<br />

introducing a circadian pattern in<strong>to</strong> the cycle<br />

such that the levels of FRQ proteins hit their<br />

lowest point late at night, and WC-1 (and WCC<br />

activity) late in the day (Dunlap & Loros, 2004).<br />

It should be noted that regulation is partly at<br />

the level of gene expression, partly by<br />

translation of existing mRNA molecules, and<br />

partly by protein phosphorylation/dephosphorylation.<br />

Entrainment of the rhythm, i.e. the<br />

switching on of the clock, is sensitive <strong>to</strong> blue<br />

light and the pho<strong>to</strong>-recep<strong>to</strong>r has been identified<br />

as the WC-1 component of WCC coupled with the<br />

chromophore FAD (flavin-adenine dinucleotide).<br />

12.2.4 Chae<strong>to</strong>mium (Chae<strong>to</strong>miaceae)<br />

There are over 80 species of Chae<strong>to</strong>mium (von Arx<br />

et al., 1986), many of which are cosmopolitan,<br />

growing in soil and fruiting on cellulose-rich<br />

substrata such as seeds, textiles in contact with<br />

soil, straw, sacking and dung. Wood infected by<br />

Chae<strong>to</strong>mium spp. may undergo a superficial decay<br />

known as soft rot. Wood inside buildings<br />

damaged by flooding or water used in fire<br />

control is particularly susceptible. Most species<br />

are saprotrophic and cellulolytic, but some have<br />

been isolated from human lesions. Some produce<br />

myco<strong>to</strong>xins (Ugadawa, 1984) and others have<br />

been used in biological control because of their<br />

competitive ability <strong>to</strong> colonize cereal stubble,<br />

thus displacing plant-pathogenic fungi (Dhingra<br />

et al., 2003). Many potentially valuable chemical<br />

compounds such as enzymes (cellulases, xylanases),<br />

pharmaceutical products and antifungal<br />

substances have been extracted from Chae<strong>to</strong>mium<br />

in culture. Chae<strong>to</strong>mium thermophile is thermophilic<br />

and has potential for use in composting palm<br />

oil fibre for recycling biomass (Suyan<strong>to</strong> et al.,<br />

2003).<br />

The perithecia of Chae<strong>to</strong>mium are superficial,<br />

barrel-shaped, thin-walled and, in most species,<br />

clothed with projecting, dark, stiff hairs. In<br />

C. elatum, one of the commonest species, the<br />

hairs are dicho<strong>to</strong>mously branched. In others,<br />

e.g. C. cochliodes, the body of the perithecium<br />

bears straight or slightly wavy, unbranched<br />

hairs, whilst the apex bears a group of spirally<br />

coiled hairs. The hairs are roughened or ornamented,<br />

and the type of ornamentation is an aid<br />

<strong>to</strong> identification (Guarro & Figueras, 1989). It is<br />

likely that the perithecial hairs have special<br />

functions. Jerking caused by the movement of<br />

the hairs over each other on drying may help in<br />

ascospore dispersal. Another possible function is<br />

<strong>to</strong> deter grazing of the perithecia by insects and


332 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.9 Chae<strong>to</strong>mium globosum. (a) Perithecium showing<br />

cirrhus of ascospores. (b) Asci. (c) Ascospores.<br />

other arthropods (Wicklow, 1979). When the<br />

perithecia are ripe, a column-like mass (cirrhus)<br />

of dark ascospores is extruded through the<br />

ostiole, supported by the perithecial hairs<br />

which surround it (Fig. 12.9a). The spore<br />

column results from the breakdown of the asci<br />

within the body of the perithecium, i.e. the asci<br />

do not discharge their spores violently. The<br />

young asci are cylindrical <strong>to</strong> club-shaped, but<br />

this stage is very evanescent and is only found<br />

in young perithecia (Fig. 12.9b), often before<br />

the spores have become pigmented and sometimes<br />

before the perithecium has developed an<br />

ostiole. Spores of most species are pale brown <strong>to</strong><br />

grey and lemon-shaped, with a single terminal<br />

germ pore.<br />

The development of perithecia in<br />

Chae<strong>to</strong>mium shows some variation between<br />

species (Whiteside, 1957, 1961). The ascogonia<br />

are coiled and lack antheridia. Investing hyphae<br />

arise from the ascogonial stalk or from vegetative<br />

cells surrounding the ascogonium.<br />

Perithecial hairs develop early from the external<br />

cell layer. The centrum is at first filled with<br />

hyaline pseudoparenchyma. At the apex of the<br />

perithecium, certain of these cells become<br />

meristematic and give rise <strong>to</strong> the elongate<br />

periphyses which line the inside of the ostiole.<br />

Ascogenous hyphae develop from the ascogonium<br />

at the base of the centrum and, at about<br />

this time, the surrounding pseudoparenchyma<br />

cells of the centrum deliquesce <strong>to</strong> form a cavity.<br />

In some species croziers develop, but in others<br />

they are absent. Paraphyses of two types have<br />

been described. In C. brasiliense, lateral paraphyses<br />

arise from the pseudoparenchyma cells<br />

outside the hymenium, whilst hymenial paraphyses<br />

have been illustrated in C. globosum<br />

(Whiteside, 1961). The paraphyses are evanescent<br />

and disappear as the asci mature.<br />

Most species of Chae<strong>to</strong>mium are homothallic,<br />

but a few, e.g. C. cochliodes, are heterothallic.<br />

Both homo- and heterothallic strains of C. elatum<br />

have been reported. Conidial states are rare in<br />

Chae<strong>to</strong>mium, but simple phialides and phialospores<br />

occur in C. elatum and C. globosum, whilst<br />

C. piluliferum forms both phialospores and<br />

globose thalloconidia of the Botryotrichum type<br />

(Daniels, 1961).<br />

12.3 Xylariales<br />

The order Xylariales is probably polyphyletic and<br />

comprises about 800 species in 8 families (Kirk<br />

et al., 2001). Members of this group produce dark<br />

perithecial stromata with asci which have an<br />

iodine-positive apical apparatus. Ascospores are<br />

typically melanized. Xylariales are saprotrophs<br />

or plant pathogens and are associated especially<br />

with the bark and wood of trees.<br />

We shall only consider one family, the<br />

Xylariaceae, which is by far the most important.<br />

The dark-coloured ascospores are generally<br />

smooth-walled and inaequilateral (one face<br />

being more strongly curved than the other),<br />

with a hyaline germ slit. Anamorphs are sometimes<br />

also stromatic, producing hyaline or<br />

lightly pigmented conidia holoblastically from<br />

a sympodially proliferating conidiogenous<br />

region, the sympodula (Fig. 12.13). They have


XYLARIALES<br />

333<br />

been placed in several anamorph genera, including<br />

Geniculosporium and Nodulisporium (Rogers,<br />

1979, 2000; Whalley, 1996). Some develop as a<br />

thin covering of the perithecial stroma. The<br />

conidia are dry and are dispersed by air currents,<br />

rain splash or insects.<br />

There are about 40 genera in the Xylariales.<br />

Most species (e.g. Xylaria, Hypoxylon, Daldinia) are<br />

hemi-saprotrophic or saprotrophic, fruiting on<br />

wood. These forms are ligninolytic and cause<br />

white-rot of their substrate (Rayner & Boddy,<br />

1988; Ea<strong>to</strong>n & Hale, 1993). The limits of the<br />

mycelial colonies in decaying wood are demarcated<br />

by black zone lines (pseudosclerotial<br />

plates) made up of brown bladder-like fungal<br />

cells which fill the wood tissue (Campbell, 1933).<br />

Other genera fruit on herbivore dung, e.g.<br />

Hypocopra, Poronia, Podosordaria and Wawelia.<br />

Some are serious plant pathogens; for example,<br />

Hypoxylon mammatum (H. pruinatum) causes<br />

canker of aspen (Populus tremuloides and other<br />

Populus species) in North America (Manion &<br />

Griffin, 1986), Biscogniauxia mediterranea causes<br />

canker of cork oak (Quercus suber), and B.<br />

nummularia causes strip canker of beech (Fagus<br />

sylvatica). Kretzschmaria (Ustulina) deusta causes a<br />

fatal butt rot of beech, elm (Ulmus) and horse<br />

chestnut (Aesculus hippocastanum), whereas<br />

Rosellinia necatrix is a plurivorous root pathogen<br />

known <strong>to</strong> attack over 130 plant species. Members<br />

of the Xylariaceae have been isolated as endophytes<br />

(symp<strong>to</strong>mless symbionts or commensals)<br />

from a range of plants on which they do not fruit<br />

so that their host range and distribution may<br />

extend far beyond that which would have been<br />

inferred from the occurrence of their ascocarps<br />

(Petrini & Petrini, 1985; Petrini et al., 1995).<br />

Although some species are plurivorous, others<br />

have a restricted host range, e.g. H. fragiforme<br />

which generally fruits on dead beech twigs<br />

(Fagus) and Xylaria carpophila which fruits only<br />

on fallen beech cupules.<br />

The development of perithecia in the family<br />

conforms <strong>to</strong> what Luttrell (1951) has termed the<br />

‘XyIaria’ type:<br />

The ascogonia are produced free upon the mycelium<br />

or more commonly within a stroma. Branches<br />

from the stalk cells of the ascogonium or from<br />

neighbouring vegetative hyphae surround the<br />

ascogonium and form the perithecial wall. Hyphal<br />

branches with free tips (paraphyses) grow upward<br />

and inward from the inner surface of the wall over<br />

the base and sides of the perithecium. Pressure<br />

exerted by the growth of opposed paraphyses<br />

expands the perithecium and creates a central<br />

cavity. The perithecium becomes pyriform as a result<br />

of growth of hyphae in the apical region of the wall<br />

<strong>to</strong> form a neck. The layer of inward-growing hyphae<br />

is continuous up the sides and in<strong>to</strong> the perithecial<br />

neck. Growth of these hyphae within the neck<br />

produces a schizogenous ostiole lined with free<br />

hyphal tips (periphyses). The ascogonium produces<br />

ascogenous hyphae which typically grow out along<br />

the inner wall over the base and sides of the<br />

perithecium. Asci derived from the ascogenous<br />

hyphae grow among the paraphyses <strong>to</strong> form a<br />

continuous hymenium of asci and more or less<br />

persistent paraphyses lining the perithecial cavity. In<br />

some forms the paraphyses are evanescent, and the<br />

ascogenous hyphae form a plexus in the base of the<br />

perithecium. The asci then arise in a single<br />

aparaphysate cluster.<br />

The cy<strong>to</strong>logy of ascus development in Xylaria<br />

(Beckett & Crawford, 1973; Rogers, 1975a) and in<br />

the related genus Hypoxylon (Rogers, 1965, 1975b)<br />

follows the usual pattern in most cases. The<br />

ascospores may be uninucleate, binucleate,<br />

or occasionally multinucleate (Rogers, 1979).<br />

In X. polymorpha and H. serpens, immature<br />

ascospores are divided by a septum near the<br />

base which cuts off a small appendage. The<br />

appendage disappears in the mature ascospore,<br />

leaving a truncate base. The ascospore wall of<br />

Daldinia concentrica, as seen with the transmission<br />

electron microscope, consists of five<br />

recognizable layers numbered progressively<br />

from the outside inwards as wall layers<br />

W1 W5 (Beckett, 1976a). The thin, non-pigmented<br />

outer layer W1 acts as a sheath which<br />

completely encloses the spore. Before germination<br />

this layer is sloughed off (see below). The<br />

wall of the mature ascospore is dark, probably<br />

due <strong>to</strong> the pigment melanin located in an inner<br />

wall layer (W4). There is a hyaline germ slit<br />

in W4 running the length of the ascospore<br />

(Fig. 12.10a). Germ slits are microfibrillar in<br />

construction but their structure varies within<br />

the family (Beckett, 1979a,b). The germ slits


334 HYMENOASCOMYCETES: PYRENOMYCETES<br />

create a weak point in the ascospore wall,<br />

causing the spore <strong>to</strong> gape wide open as it swells<br />

and permitting the germ tubes <strong>to</strong> emerge<br />

(Beckett, 1976b; Chapela et al., 1990; Read &<br />

Beckett, 1996). The ascus tip contains a cylindrical<br />

apical apparatus which is amyloid (i.e. it<br />

stains bright blue with iodine). The apical<br />

apparatus is pierced by a narrow pore and is<br />

everted as the ascus explodes (Greenhalgh &<br />

Evans, 1967; Beckett & Crawford, 1973).<br />

12.3.1 Daldinia<br />

There are about 13 species of Daldinia worldwide<br />

(Ju et al., 1997), 5 of which grow in Northern<br />

Europe (Johannesson et al., 2000). They form<br />

concentrically zoned perithecial stromata on<br />

wood. The best known is D. concentrica, with<br />

stromata mostly on ash (Fraxinus excelsior) but<br />

occasionally on other hosts. Daldinia loculata<br />

(D. vernicosa) and D. fissa form stromata on<br />

charred branches of birch (Betula), gorse (Ulex)<br />

and some other hosts following fire.<br />

Daldinia concentrica colonizes the living<br />

branches of ash as a symp<strong>to</strong>mless endophyte<br />

but can continue <strong>to</strong> grow saprotrophically,<br />

fruiting on dead branches and trunks. Recently<br />

infected wood (calico wood) has a dark speckled<br />

appearance caused by the presence of a darkcoloured<br />

mycelium in the vessels of the spring<br />

wood. Boddy et al. (1985) isolated D. concentrica<br />

from attached branches of ash and regarded it as<br />

a primary colonizer. Only a small number of<br />

individual mycelia were isolated from any one<br />

branch, but often these occupied extensive<br />

volumes of wood. Individual mycelia are<br />

detected by their reactions in culture <strong>to</strong> other<br />

individuals of the same species. Identical mycelia<br />

intermingle freely, showing no obvious reaction,<br />

but mycelia of different genotype show vegetative<br />

incompatibility when confronted. Such<br />

observations have led <strong>to</strong> the suggestion that<br />

the colonization of attached branches is by a<br />

process of latent invasion (Boddy & Rayner, 1983)<br />

whereby the fungus might be distributed in the<br />

sap stream of the living tree host as mycelial<br />

fragments or spores. It is likely that D. concentrica<br />

is heterothallic (Sharland & Rayner, 1986) like<br />

D. loculata which grows on fire-scorched birch<br />

branches (Johannesson et al., 2001). In this<br />

fungus, the Nodulisporium conidial stroma develops<br />

in spring beneath the birch bark before the<br />

perithecial stroma emerges. Pyrophilous (fireloving)<br />

insects feed on the conidia and also<br />

disperse them. Conidia of different mating types<br />

and different genotypes may be spread in this<br />

way, and since some perithecial stromata are<br />

known <strong>to</strong> contain several genetically distinct<br />

perithecia, multiple mating events are possibly<br />

involved.<br />

Fig12.10 (a,b) Xylaria hypoxylon. (a) Ascus.The ascus tip <strong>to</strong> the<br />

right has been stained with iodine <strong>to</strong> reveal the apical<br />

apparatus. (b) Conidiophores. (c,d) Daldinia concentrica.<br />

(c) Ascogenous hypha (after Ingold,1954b).The numbers<br />

represent successive asci working backwards from the apex.<br />

(d) Nodulisporium-type conidiophore.


XYLARIALES<br />

335<br />

Daldinia concentrica forms large (5 10 cm<br />

diameter) hemispherical purplish-brown annual<br />

stromata called ‘cramp balls’ or ‘King Alfred’s<br />

cakes’. They contain ripe asci between May and<br />

Oc<strong>to</strong>ber. In cross section (Plate 5a), the stromata<br />

show a concentric zonation of alternating light<br />

and dark bands. The surface of young stromata<br />

may be covered with a pale fawn powdery mass<br />

of conidia. The conidia are dry and ovoid in<br />

shape, developing successively at the tips of<br />

branched conidiophores by the outgrowth of<br />

the wall and, when detached, leave a small scar<br />

(Fig. 12.10d). Conidia of this type have been<br />

named Nodulisporium tulasnei. Perithecia develop<br />

in the outer layers of the stroma. The perithecial<br />

wall is lined by ascogenous hyphae which are<br />

unusual in that there is often a considerable<br />

distance separating successive asci (Ingold,<br />

1954b) (Fig. 12.10c). The stroma of Daldinia<br />

apparently functions as a water reservoir and<br />

detached stromata will continue <strong>to</strong> discharge<br />

ascospores for about 3 weeks even if placed in a<br />

desicca<strong>to</strong>r (Ingold, 1946). Spore discharge is<br />

nocturnal and the rhythm of spore discharge is<br />

maintained for several days if detached stromata<br />

are kept in continuous dark. In continuous light,<br />

periodic spore discharge ceases after about three<br />

days but is res<strong>to</strong>red on return <strong>to</strong> alternating light<br />

and dark (Ingold & Cox, 1955). The output<br />

of spores from a single stroma of average size<br />

is about 10 million a night. The ability of<br />

perithecial stromata <strong>to</strong> s<strong>to</strong>re water enables the<br />

fungus <strong>to</strong> continue sporulating on dry branches.<br />

Mycelial growth is also possible at lower water<br />

potentials ( 10 MPa) than by competing fungi<br />

(Boddy et al., 1985).<br />

12.3.2 Xylaria<br />

There are over 100 species of Xylaria, most of<br />

which are lignicolous, but some are endophytic<br />

and others grow on fallen fruits (Whalley, 1985,<br />

1987). One of the best known is Xylaria hypoxylon,<br />

the candle-snuff fungus. Stromata are common<br />

on stumps and fallen branches of deciduous<br />

trees. As in most Xylariaceae growing on wood,<br />

the boundaries of individual mycelia within<br />

infected tissues are visible as conspicuous black<br />

demarcation lines. The stromata are branched<br />

and cylindrical or flattened. At the upper end,<br />

the stroma is covered by a white powdery mass<br />

of conidia (Figs. 12.10b, 12.11a). Perithecia<br />

develop later at the base of the stroma and are<br />

visible externally as swellings at the surface<br />

(Fig. 12.11b). The apical apparatus of the ascus is<br />

visible even in immature asci after staining in<br />

iodine as a bright blue cylindrical collar pierced<br />

by a narrow pore (Fig. 12.10a). Xylaria polymorpha<br />

(‘dead men’s fingers’) fruits in late summer and<br />

autumn at the base of old tree stumps. The<br />

stromata are swollen, finger-like and clustered<br />

(Plate 5b). The surface is at first covered by an<br />

inconspicuous conidial layer, but eventually<br />

perithecia develop beneath the surface of the<br />

whole stroma, and are not restricted <strong>to</strong> the basal<br />

region as in X. hypoxylon. Both species are active<br />

wood-rotting fungi causing decay of the whiterot<br />

type. Other common species with a more<br />

restricted host range are X. longipes on fallen<br />

branches of sycamore (Acer pseudoplatanus) and<br />

X. carpophila on fallen cupules of beech (Fagus<br />

sylvatica).<br />

12.3.3 Hypoxylon<br />

This is a large genus of over 120 species (Whalley<br />

& Greenhalgh, 1973; Ju & Rogers, 1996) forming<br />

stromata which are often hemispherical or<br />

sometimes flattened on the surface of wood<br />

and bark. Some show a preference for a particular<br />

host. Common species are H. fragiforme<br />

almost confined <strong>to</strong> branches and trunks of<br />

Fagus (Fig. 12.12a), H. multiforme on Betula<br />

(Fig. 12.12b), and H. rubiginosum which forms<br />

flat stromata on decorticated wood of Fraxinus.<br />

The young stromata of all these species<br />

bear a conidial felt of the Nodulisporium or<br />

Geniculosporium type (Fig. 12.13; Chesters &<br />

Greenhalgh, 1964). Most species show nocturnal<br />

spore discharge (Walkey & Harvey, 1966b).<br />

Freshly cut lengths of healthy beech branches<br />

incubated under water-saturated conditions<br />

show no evidence of the presence of H. fragiforme,<br />

but if similar sections are incubated under<br />

conditions in which the branches are allowed<br />

<strong>to</strong> dry, characteristic patches of stained or<br />

discoloured wood become apparent within<br />

21 days (Chapela & Boddy, 1988a,b). Isolations<br />

from such areas produce several genetically


336 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.11 Xylaria hypoxylon. (a) Conidial<br />

stroma.Conidia are borne on the white<br />

tips of the branches. (b) Perithecial<br />

stromata which have developed at the<br />

base of the old conidial stroma.The<br />

knobbly swellings are the perithecia.<br />

distinct mycelia of H. fragiforme, indicating that<br />

the infections are derived from a number of<br />

separate ascospores. The fungus is present in<br />

apparently healthy sapwood as a number of<br />

inconspicuous pockets of ‘inoculum units’ from<br />

which mycelial growth is held in check by the<br />

high water content of the wood. On drying,<br />

these infections can spread rapidly. This phenomenon<br />

is another example of latent invasion<br />

(Boddy & Rayner, 1983). Chapela (1989) used the<br />

term ‘xylotropic endophyte’ for fungi such as<br />

H. fragiforme which occur within living trees and<br />

have the capacity <strong>to</strong> extend in<strong>to</strong> secondary xylem<br />

upon drying of the wood.<br />

Ascospore eclosion<br />

The results of experimental studies on ascospore<br />

germination in H. fragiforme provide a partial<br />

explanation of its host specificity. Although<br />

known <strong>to</strong> fruit on several other woody hosts,<br />

perithecial stromata are most regularly associated<br />

with drying branches and trunks of<br />

European beech (Fagus sylvatica) and American<br />

beech (F. grandiflora). Ascospores suspended in<br />

extracts from living F. sylvatica twigs react by a<br />

dramatic germination process termed eclosion,<br />

an en<strong>to</strong>mological term for the emergence of<br />

a pupa from its case or a larva from an egg<br />

(Chapela et al., 1990, 1993). In H. fragiforme this<br />

appears <strong>to</strong> be a host-specific recognition system.<br />

Exposure <strong>to</strong> aqueous extracts of beech twigs for<br />

a period of about 10 min activates the eclosion<br />

mechanism, which is a two-stage process. In the<br />

first stage the spore swells slightly and its germ<br />

slit widens until the pigmented spore wall opens<br />

abruptly along the germ slit, forcing the outer<br />

transparent sheath (corresponding <strong>to</strong> the W1<br />

layer of D. concentrica) <strong>to</strong> crack open transversely.<br />

This stage is a millisecond event. The second<br />

stage, lasting about 10 s, involves the widening<br />

of the germ slit up <strong>to</strong> about 7 mm, and expansion<br />

of the coloured wall of the ascospore. This<br />

forces the transparent outer sheath away from<br />

the rest of the spore, which then escapes and is<br />

free <strong>to</strong> germinate by the formation of a germ<br />

tube. Eclosion is readily observed with the light<br />

microscope, as summarized in Fig. 12.14 (see<br />

Webster & Weber, 2004).<br />

The triggers which stimulate eclosion have<br />

been identified as two monolignol glucosides,<br />

Z-syringin and Z-isoconiferin, which are mobile<br />

transport intermediates of lignin biosynthesis.<br />

Both induce eclosion at micromolar concentrations.<br />

However, the presence of these germination<br />

activa<strong>to</strong>rs in beech extracts cannot fully<br />

explain the host specificity of H. fragiforme<br />

because eclosion is induced by extracts from a<br />

range of trees which do not normally support<br />

fruiting of this fungus, such as Abies, Corylus and<br />

Populus (Chapela et al., 1991).


HYPOCREALES<br />

337<br />

Fig12.13<br />

serpens.<br />

The conidial (Geniculosporium)stateofHypoxylon<br />

Fig12.12 Hypoxylon. (a) Perithecial stromata of H. fragiforme<br />

on beech (Fagus sylvatica).One stroma has been broken open<br />

<strong>to</strong> show the perithecia embedded in the outer layer.<br />

(b) Perithecial stromata of H. multiforme on birch (Betula<br />

pendula).<br />

12.4 Hypocreales<br />

The Hypocreales are a large group of fungi,<br />

although estimates vary as <strong>to</strong> the number of taxa<br />

contained in it. Kirk et al. (2001) included 117<br />

genera and 654 species, whilst Rossman (1996)<br />

stated that there may be 2000 5000 holomorphic<br />

taxa. Some are known only as anamorphs.<br />

Hypocreales are characterized by pale or brightly<br />

coloured perithecia (or rarely cleis<strong>to</strong>thecia)<br />

which may be single or borne on or embedded<br />

in a fleshy stroma. The asci are unitunicate, with<br />

or without a well-defined apical apparatus.<br />

Perithecial development conforms <strong>to</strong> the<br />

‘Nectria’ type of Luttrell (1951). The ascogonia,<br />

which are formed within a stroma, become<br />

surrounded by concentric layers of vegetative<br />

hyphae which form a true perithecial wall. The<br />

cells of the inner wall layer in the apical region<br />

of the young perithecium produce a palisade of<br />

inward-growing hyphal branches. These hyphal<br />

branches grow downward <strong>to</strong> form a vertically<br />

arranged mass of hyphae with free ends termed<br />

apical paraphyses (Luttrell, 1965). Pressure<br />

exerted by the elongation of the apical paraphyses,<br />

accompanied by expansion of the wall,<br />

creates a central cavity within the perithecium.<br />

The free tips of the apical paraphyses ultimately<br />

push in<strong>to</strong> the lower portion of the wall so that<br />

they become attached at both the <strong>to</strong>p and<br />

bot<strong>to</strong>m of the perithecial cavity. Ascogenous<br />

hyphae arising from the ascogonium spread out<br />

across the floor and sides of the cavity and<br />

produce asci by means of croziers. The asci grow<br />

upward among the apical paraphyses and form


338 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.14 The course of events during eclosion of Hypoxylon fragiforme ascospores. (a,b) Dormant spores as seen in median (a) and<br />

surface view (b), the latter showing the germ slits (arrowheads). (c e) Rupture of wall layer W1along a longitudinal seam, gaping of<br />

germ fissure of ascospore, and escape of the spore from the perispore shell formed by wall layer W1.The sequence (c e) can be<br />

observed some10 min after exposure of ascospores <strong>to</strong> the beech extract. It occurs abruptly, taking about 3 s. (f) Further distension<br />

of valves composed of wall layers W2 W4, 30 min after addition of beech twig extract.The perispores are still visible as ghosts.<br />

(g) Formation of a germ tube some 8 h after addition of beech twig extract. All images <strong>to</strong> same scale. Reprinted from Webster<br />

and Weber (2004), with permission from Elsevier.<br />

a concave layer lining the inner surface of the<br />

wall in the basal region of the perithecium.<br />

At the upper end of the perithecium a schizogenous<br />

ostiole develops in the wall. It is lined<br />

with periphyses, i.e. hyphae with free apices<br />

which are attached at their bases <strong>to</strong> the inner<br />

wall of the neck. At maturity, the perithecia may<br />

protrude from the stroma and appear <strong>to</strong> be<br />

seated on its surface. Perithecial development<br />

of this type has been described for Nectria<br />

(Strickmann & Chadefaud, 1961; Hanlin, 1971),<br />

and for Hypocrea (Canham, 1969).<br />

There is an exceptionally wide range of<br />

anamorphs (Rossman, 2000; Seifert & Gams,<br />

2001). In most cases conidiogenous cells are<br />

phialides, and these may be terminal or lateral,<br />

single or grouped in synnemata, sporodochia or,<br />

more rarely, pycnidia. However, other<br />

developmental types of conidia may occur and<br />

in many species there are synanamorphs. The<br />

conidia are generally light-coloured and<br />

produced in slimy masses. The names of some<br />

anamorph genera related <strong>to</strong> Hypocreales are<br />

listed in Table 12.2. There is a tendency in<br />

modern classification, with support of molecular<br />

evidence, <strong>to</strong> link anamorph states with teleomorph<br />

genera even if they have, as yet, no<br />

proven connection from pure culture studies<br />

(Rossman, 2000).<br />

Many Hypocreales are saprotrophs active in<br />

the decay of plant substrata above ground, in soil<br />

or fresh water. Others are serious plant pathogens<br />

(e.g. Nectria and Fusarium spp.) or mycoparasites,<br />

especially of agaric, bolete or polypore<br />

basidiocarps (e.g. Hypomyces, Apiocrea and<br />

Hypocrea; Põldmaa, 2000) but also of cultivated


HYPOCREALES<br />

339<br />

mushroom mycelium. Some are lichenicolous.<br />

The ability <strong>to</strong> parasitize other fungi has been<br />

employed for the biological control of fungal<br />

pathogens of plants, using species of Trichoderma,<br />

Gliocladium and Clonostachys (see below).<br />

Several antifungal compounds, e.g. glio<strong>to</strong>xin<br />

or viridin, are produced as secondary metabolites<br />

by anamorphs of Hypocrea spp. (Trichoderma and<br />

Gliocladium; see below). Other pharmacologically<br />

important secondary metabolites synthesized<br />

by anamorphic Hypocreales include the antibacterial<br />

antibiotic cephalosporin from Acremonium<br />

spp. The gibberellins are growth hormones<br />

ubiqui<strong>to</strong>us in all higher plants, but they were<br />

first detected as a secondary metabolite<br />

produced by Gibberella fujikuroi (anamorph<br />

Fusarium moniliforme). This species causes the<br />

‘foolish seedling disease’ in rice, in which<br />

seedlings show excessive stem elongation and<br />

eventually keel over due <strong>to</strong> the production of<br />

gibberellic acid by the infecting fungus<br />

(Tudzynski, 1997). Myco<strong>to</strong>xins poisonous <strong>to</strong><br />

farm animals and humans are produced by<br />

some species of Fusarium (see below) and other<br />

species have been reported as human pathogens.<br />

The edible fungus food ‘Quorn’ is<br />

Table12.2. Some teleomorph anamorph<br />

connections in the Hypocreales.<br />

Teleomorph genus<br />

Nectria sensu la<strong>to</strong><br />

Bionectria<br />

Calonectria<br />

Gibberella<br />

Hypocrea<br />

Sphaerostilbella<br />

Hypomyces<br />

Apiocrea<br />

Melanopsamma<br />

Anamorph genus<br />

Tubercularia, Fusarium,<br />

Cylindrocarpon,<br />

Verticillium, Heliscus,<br />

Flagellospora<br />

Clonostachys<br />

Cylindrocladium<br />

Fusarium<br />

Trichoderma,Gliocladium,<br />

Acremonium-like<br />

Gliocladium<br />

Mycogone,Cladobotryum<br />

Sepedonium<br />

Stachybotrys<br />

manufactured by large-scale fermentation of<br />

F. venenatum (previously identified as F. graminearum)<br />

(Trinci, 1991; Moore & Chiu, 2001).<br />

Cellulase production by Trichoderma reesei is<br />

exploited commercially.<br />

General accounts of the Hypocreales have<br />

been given by Rogerson (1970) with keys <strong>to</strong><br />

genera, and by Rossman (1996). The taxonomic<br />

treatment, i.e. the division in<strong>to</strong> orders and<br />

families, varies between different authors, some<br />

placing Claviceps and its allies as a family<br />

(Clavicipitaceae) of the Hypocreales, whilst<br />

others, e.g. M. Barr (2001), placed them in a<br />

separate order (Clavicipitales), a disposition<br />

followed in this book. However, all are agreed<br />

that the two groups are closely related, a view<br />

based on morphological and molecular evidence<br />

(Spatafora & Blackwell, 1993). Samuels and<br />

Blackwell (2001) included four families in the<br />

Hypocreales, i.e. Hypocreaceae, Nectriaceae,<br />

Bionectriaceae and Niessliaceae, but we shall<br />

consider representatives only of the first two.<br />

12.4.1 Hypocrea (Hypocreaceae)<br />

Species of Hypocrea, of which about 100 are<br />

known, usually fruit on decaying wood or<br />

occasionally on herbaceous plant material, forming<br />

brightly coloured fleshy perithecial stromata,<br />

with perithecia embedded in the outer<br />

layers. The thin-walled asci contain 8 two-celled<br />

ascospores, and in many species the spores<br />

separate in<strong>to</strong> 2 part-spores before ascus<br />

discharge, so that 16 part-ascospores are released<br />

(Fig. 12.15c). The ascospores may be colourless<br />

(hyaline) or green in colour. Hypocrea pulvinata<br />

forms bright yellow stromata on the underside<br />

of dead overwintered basidiocarps of Pip<strong>to</strong>porus<br />

betulinus, the birch polypore. It is possible that<br />

this fungus grows parasitically on the basidiocarp.<br />

The ascospores are often visible as white<br />

tendrils issuing from the ostioles of the perithecia<br />

(Plate 5c). In culture, conidia are formed in<br />

sticky masses at the tip of single phialides<br />

(Fig. 12.15b; Rifai & Webster, 1966). Conidia of<br />

this type are described as Acremonium-like.<br />

Some species of Hypocrea have conidia of the<br />

Trichoderma type, in which whorls of phialides<br />

give rise <strong>to</strong> separate, sticky green or white


340 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.15 Hypocrea pulvinata.(a)L.S.<br />

lower part of fruit body of Pip<strong>to</strong>porus<br />

betulinus showing perithecial stromata<br />

of Hypocrea in section (see also<br />

Plate 5c). (b) Acremonium-like conidia<br />

produced from upright phialides.<br />

(c) Asci and ascospores. Note how<br />

the two-celled ascospores may<br />

break up in<strong>to</strong> unicellular part-spores.<br />

spore masses. Detailed descriptions of Hypocrea<br />

Trichoderma connections have been given by<br />

Chaverri and Samuels (2003). Hypocrea rufa<br />

forms conidia referable <strong>to</strong> T. viride (Fig. 12.16d),<br />

a species with globose, warty conidia. Conidia of<br />

the Gliocladium type, in which conidia derived<br />

from individual phialides coalesce in a single<br />

slimy mass, are found in cultures of H. gelatinosa<br />

(Figs. 12.16a,b; Webster, 1964). The distinction<br />

between Trichoderma and Gliocladium anamorphic<br />

states can be difficult and some species, e.g. G.<br />

virens, have been transferred <strong>to</strong> Trichoderma as T.<br />

virens (Fig. 12.17a), the anamorph of H. virens<br />

(Chaverri et al., 2001).<br />

The biology and taxonomy of Trichoderma<br />

have been reviewed by Samuels (1996, 2006),<br />

ecological aspects by Klein and Eveleigh (1998),<br />

and keys <strong>to</strong> identification have been given by<br />

Gams and Bissett (1998). It is a large genus and<br />

species identification is difficult using morphological<br />

criteria. DNA-based and biochemical<br />

techniques are now widely used <strong>to</strong> support<br />

identification. Such studies indicate that<br />

genus is a monophyletic group within the<br />

Hypocreaceae which evolved about 110 million<br />

years ago (Kullnig-Gradinger et al., 2002).<br />

Trichoderma spp. are cosmopolitan in soil and<br />

on decaying woody substrata. On soil isolation<br />

plates, Trichoderma spp. are often the most<br />

rapidly growing and dominant fungi, smothering<br />

the plates and forming clusters of sticky<br />

green phialoconidia within a few days. Certain<br />

species, especially T. harzianum and T. virens, have<br />

been used in the biological control of soil-borne<br />

fungal plant pathogens such as Rhizoc<strong>to</strong>nia solani<br />

and Pythium ultimum (Papavizas, 1985; Chet, 1987;<br />

Lumsden, 1992; Tang et al., 2001). Antagonism <strong>to</strong><br />

the fungal pathogen may be associated with<br />

coiling of Trichoderma around the host hypha,<br />

followed by penetration and parasitism<br />

(Figs. 12.17b d). Volatile and non-volatile<br />

antifungal antibiotics may also be produced.<br />

Several species of Trichoderma are troublesome<br />

contaminants and parasites of mycelia and<br />

basidiocarps of cultivated mushrooms in<br />

Europe and North America (Castle et al., 1998;<br />

Samuels et al., 2002).<br />

Trichoderma reesei is used commercially for<br />

cellulase production, secreting 20 g or more of<br />

cellulase l 1 culture fluid (Durand et al., 1988;<br />

Kubicek et al., 1990). Until recently, this interesting<br />

species had been isolated only once, during<br />

the Second World War on the Solomon Islands<br />

from canvas in contact with soil. Molecular<br />

studies have revealed that T. reesei is the<br />

anamorph of Hypocrea jecorina, an uncommon<br />

tropical species (Kuhls et al., 1996; Lieckfeldt<br />

et al., 2000).


HYPOCREALES<br />

341<br />

Fig12.16 Conidiophores of<br />

Hypocrea spp. (a) Gliocladium-type<br />

conidiophores of H. gelatinosa.<br />

(b) Details of phialides of<br />

H. gelatinosa.(c)Trichoderma viride<br />

conidial state of H. rufa. (d) Detail<br />

of phialides of H. rufa. (a,c) <strong>to</strong> same<br />

scale; (b,d) <strong>to</strong> same scale.<br />

The mating behaviour of Hypocrea spp. is<br />

poorly unders<strong>to</strong>od because perithecial development<br />

in culture occurs only rarely, except for the<br />

tropical species H. jecorina. This species shows<br />

bipolar heterothallism, half of its ascospores<br />

being of one and half of the opposite mating<br />

type (Lieckfeldt et al., 2000). Another tropical<br />

species, H. poronioidea, also forms perithecia in<br />

culture but the genetic basis of sexual reproduction<br />

is not clear. Cultures from eight partascospores<br />

in each of several asci were self-fertile<br />

(i.e. homothallic), whilst the other eight partascospores<br />

produced only conidia and were selfsterile.<br />

Another unusual feature of this fungus<br />

is that, in addition <strong>to</strong> having a Trichoderma<br />

anamorph, it has an Acremonium-like synanamorph.<br />

The Acremonium-like conidia can germinate,<br />

i.e. they have an asexual function, but it is<br />

possible that they also have a spermatial role<br />

(Samuels & Lodge, 1996).<br />

12.4.2 Nectria (Nectriaceae)<br />

The concept of the genus Nectria has changed in<br />

recent years. Previously considered <strong>to</strong> include<br />

about 200 species, the genus has now been<br />

narrowed <strong>to</strong> about 30, centred around<br />

N. cinnabarina (Rossman, 1983; Rossman et al.,<br />

1999). Perithecia of Nectria are common on twigs<br />

and branches of woody hosts. Many are saprotrophic<br />

or weakly parasitic but some cause<br />

economically important diseases, e.g. N. coccinea<br />

causes bark disease of beech (Fagus sylvatica),<br />

whereas N. galligena infects apple and pear trees.<br />

Although there are some reports that N. galligena<br />

may have a prolonged latent (endophytic) phase,<br />

the fungus acts as a wound pathogen in most


342 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.17 Mycoparasitism byTrichoderma spp. (a) Tip of conidiophore of T. virens with phialides. (b) Contact of a hypha of T. harzianum<br />

with one of Rhizoc<strong>to</strong>nia cerealis. (c) Coiling of T. harzianum around a hypha of R. cerealis. (d) Intrahyphal growth of T. virens in a hypha<br />

of Fusarium sp.<br />

situations, entering through pruning wounds,<br />

cracks, leaf scars and sites of branch breakage.<br />

The colonization of the bark and cambium<br />

tissues results in a necrotic lesion, often with<br />

the dead heartwood visible in the centre, and<br />

surrounded by raised callus-like bark tissue<br />

where the tree launches a defence response.<br />

This type of symp<strong>to</strong>m is called a canker<br />

(Fig. 12.18). Conidia and, later, perithecia are<br />

produced in the bark of cankers. Girdling<br />

results if a canker surrounds an entire branch,<br />

and this leads <strong>to</strong> the death of all shoot tissue<br />

distal <strong>to</strong> the lesion.<br />

The conidial states of Nectria include some<br />

terrestrial species of Fusarium, Cylindrocarpon and<br />

Tubercularia, whilst in freshwater streams they<br />

include the aquatic hyphomycetes Cylindrocarpon,<br />

Flagellospora and Heliscus (Section 25.2; Webster,<br />

1992). Some species have synanamorphic states<br />

with conidia of distinctive size and shape, e.g.<br />

macro- and micro-conidia (e.g. N. haema<strong>to</strong>cocca;<br />

Fig. 12.19).<br />

Nectria cinnabarina, the cause of coral spot, is<br />

common on freshly cut twigs of hardwood trees<br />

and shrubs, but may occasionally be a wound<br />

parasite on these hosts. The name coral spot<br />

refers <strong>to</strong> the pale pink conidial pustules, about<br />

1 2 mm in diameter, which burst through the<br />

bark (Plate 5d). Before the connection with<br />

Nectria was unders<strong>to</strong>od, these conidial pustules<br />

had been named Tubercularia vulgaris. They<br />

consist of a column of pseudoparenchyma bearing<br />

a dense tuft of conidiophores. These are long<br />

slender hyphae producing intercalary phialides<br />

at intervals along their length (Figs. 12.20b,d).<br />

Conidial pustules of this type are termed<br />

sporodochia. The conidia are sticky and form a<br />

slimy mass at the surface of the sporodochium.<br />

They are dispersed very effectively by rain splash.<br />

Around the base of the old conidial pustule<br />

perithecia arise (Plate 5d), and eventually the<br />

pustule may bear perithecia over its entire<br />

surface. Perithecial pustules develop in damp<br />

conditions in late summer and autumn and are


HYPOCREALES<br />

343<br />

Fig12.18 Canker caused by Nectria galligena<br />

on the twig of an old apple tree.<br />

readily distinguished from conidial pustules by<br />

their bright red colour and their granular<br />

appearance. The pustules are regarded as perithecial<br />

stromata, bearing as many as 30 perithecia.<br />

Ripe perithecia contain numerous<br />

club-shaped asci, each with eight two-celled<br />

hyaline ascospores (Fig. 12.20c). These are somewhat<br />

unusual in being multinucleate (El-Ani,<br />

1971). There is no obvious apical apparatus <strong>to</strong> the<br />

ascus (Strickmann & Chadefaud, 1961).<br />

Perithecial development begins by the formation<br />

of ascogonial primordia beneath the surface<br />

of the conidial pustule. The details of development<br />

are as described on p. 337. An important<br />

feature is the development of apical paraphyses<br />

which grow downwards from the upper part of<br />

the perithecial cavity. As the asci develop from<br />

ascogenous hyphae lining the base of the<br />

perithecial cavity, they grow upwards through<br />

the mass of apical paraphyses, which are difficult<br />

<strong>to</strong> find in mature perithecia (Strickmann &<br />

Chadefaud, 1961).<br />

Other species of Nectria differ from N. cinnabarina<br />

in a number of ways. In many, the perithecia<br />

are not grouped <strong>to</strong>gether on a stroma, but<br />

occur singly (e.g. in N. galligena). The asci of some<br />

species, e.g. N. mammoidea, have a well-defined<br />

apical apparatus (Fig. 12.21a).<br />

12.4.3 Fusarium (Nectriaceae)<br />

Fusarium-type conidia are known in several<br />

species of Nectria and also in the related genus<br />

Gibberella (Samuels et al., 2001). For many<br />

Fusarium species a teleomorph has not yet been<br />

found. The number of species recognized by<br />

morphological characters in pure culture varies<br />

from one authority <strong>to</strong> another. Booth (1971)<br />

included 44 species and 7 varieties; Gerlach and<br />

Nirenberg (1982) included 73 species and 26<br />

varieties, while Nelson et al. (1983) distinguished<br />

30 species. However, using molecular techniques<br />

a large number of species which cannot be<br />

distinguished by morphological characters are<br />

now being defined (O’Donnell, 1996), and a<br />

publically accessible database of diagnostic DNA<br />

sequences has been established as an aid <strong>to</strong> identification<br />

(Geiser et al., 2004). Species of Fusarium<br />

are cosmopolitan in soils from the permafrost of<br />

the Arctic <strong>to</strong> sand in the Sahara desert (Booth,<br />

1971). They are particularly common in cultivated<br />

soil and are associated with a wide range of<br />

plant diseases as indicated in Table 12.3 (Nelson<br />

et al., 1981). Some species are also known <strong>to</strong><br />

cause diseases of humans such as onychomycosis<br />

(nail infections), kera<strong>to</strong>mycosis of the cornea,<br />

ulcers, necroses, skin infections and fatal infections<br />

of internal organs, especially in immunocompromised<br />

patients (Joffe, 1986; de Hoog et al.,<br />

2000a).<br />

A characteristic feature in the asexual reproduction<br />

of Fusarium is the development from<br />

phialides of fusoid, transversely septate macroconidia<br />

with a basal contracted foot cell<br />

(Fig. 12.22b). Microconidia, also formed from<br />

phialides, develop in some species. On plant<br />

hosts macroconidia often accumulate in<br />

orange pink sporodochia (Plate 5e, Fig. 12.22a).<br />

Macroconidia may be dispersed by rain splash


344 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.19 Nectria haema<strong>to</strong>cocca.<br />

(a) Ascospores. (b) Phialides bearing<br />

macroconidia. (c) Macroconidia of the<br />

Fusarium type. (d) Phialides producing<br />

microconidia which accumulate in a drop<br />

of mucilage. (e) Microconidia.<br />

Fig12.20 Nectria cinnabarina.(a)V.S.<br />

perithecial stroma. (b) V.S. conidial<br />

stroma (sporodochium). (c) Asci.<br />

(d) Conidiophore, phialides and conidia.


HYPOCREALES<br />

345<br />

Fig12.21 Nectria mammoidea. (a) Ascus;<br />

note the apical apparatus.<br />

(b) Cylindrocarpon-type conidiophore with<br />

phialides and conidia.<br />

from such pustules (see Jenkinson & Parry,<br />

1994). In culture, greasy accumulations of macroconidia<br />

are termed pionnotes. Survival in soil is<br />

probably in the form of dormant chlamydospores<br />

which may be formed as a result of energy<br />

deprivation or in response <strong>to</strong> bacterial secretions.<br />

They are thick-walled and develop<br />

by modification of segments of the vegetative<br />

mycelium or by enlargement and modification<br />

of a segment of a macroconidium (Schippers &<br />

van Eck, 1981) (Figs. 12.22c,d). Chlamydospores<br />

are induced <strong>to</strong> germinate in the presence of root<br />

secretions (for references see Griffiths, 1974).<br />

One of the most commonly isolated species<br />

from soil is F. oxysporum, which often grows in<br />

association with roots. Many isolates are nonpathogenic<br />

but this fungus may also be a serious<br />

plant pathogen, reported from a very wide range<br />

of plant hosts. Although no sexual state is<br />

known, molecular studies show that it is monophyletic<br />

with the Gibberella Fusarium complex<br />

and can be regarded as an asexual Gibberella<br />

(Samuels et al., 2001). In this species over 80<br />

formae speciales characteristic of different host<br />

genera have been recognized, in addition <strong>to</strong> a<br />

large number of vegetative compatibility types<br />

and races (Gordon, 1993; Gordon & Martyn,<br />

1997). The large amount of genetic variation in<br />

F. oxysporum in the absence of conventional<br />

sexual reproduction is possibly explained by<br />

the demonstration of parasexual recombination<br />

even between members of different vegetative<br />

compatibility groups (Molnár et al., 1990).<br />

Fusarium wilts<br />

As can be seen from Table 12.3, formae speciales of<br />

F. oxysporum cause wilt diseases of numerous crop<br />

plants. The range of affected hosts is so wide that<br />

it is easier <strong>to</strong> mention those plants unaffected by<br />

Fusarium wilts, i.e. grasses and many tree species.<br />

Although F. oxysporum has no known sexual state,<br />

strains of this species complex show huge<br />

genetic variation which may come about by the<br />

presence of transposable elements (Daboussi &<br />

Capy, 2003) and the parasexual cycle (Teunissen<br />

et al., 2002). The variability of F. oxysporum is<br />

such that a given disease, e.g. Panama wilt of<br />

banana, is caused by several unrelated strains<br />

(K. O’Donnell et al., 1998), and that strains with<br />

different host specificities may be closely related<br />

(Hua-Van et al., 2001).<br />

Wilting is associated with the presence of<br />

fungal hyphae in the xylem vessels. Fusarium<br />

oxysporum can persist in the soil as chlamydospore<br />

inoculum for many years and infects the<br />

roots of hosts by direct penetration and


346 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Table 12.3. Some species of Fusarium causing plant diseases of economic importance. From Holliday (1998).<br />

Anamorph Teleomorph Disease caused and name of host<br />

F. culmorum not known cortical rot, foot rot and pre-emergence blight of<br />

temperate cereals; head blight of barley,<br />

wheat and rye<br />

F. oxysporum not known a wide range of wilts, yellows and foot rots<br />

F. oxysporum f. sp. apii not known celery yellows<br />

F. oxysporum f. sp. cepae not known basalrot and s<strong>to</strong>rage rot of onion<br />

F. oxysporum f. sp. cubense not known Panama wilt of banana<br />

F. oxysporum f. sp. dianthi not known wilt of carnation and pinks<br />

F. oxysporum f. sp. elaeidis not known oil palm wilt<br />

F. oxysporum f. sp. lycopersici not known <strong>to</strong>ma<strong>to</strong> wilt<br />

F. oxysporum f. sp. pisi not known pea wilt<br />

F. oxysporum f. sp. vasinfectum not known cot<strong>to</strong>n wilt<br />

F. solani Nectria<br />

haema<strong>to</strong>cocca<br />

root and collar rot of many plants; canker of<br />

woody crops<br />

F. solani f. sp. cucurbitae not known foot rot of cucumber and melon<br />

F. solani f. sp. coeruleum not known s<strong>to</strong>rage rot and dry rot of pota<strong>to</strong><br />

F. avenaceum Gibberella<br />

avenacea<br />

damping off and rootdamage <strong>to</strong> cereals, conifers<br />

andlegumes<br />

F. sulphureum G. cyanogena s<strong>to</strong>rage rot and dry rot of pota<strong>to</strong> tubers<br />

F. moniliforme G. fujikuroi diseases of many plants, e.g. banana black heart,<br />

cot<strong>to</strong>n bollrot, maize and sorghum stalk rots,<br />

rice seedling bakanae disease<br />

F. sambucinum G. pulicaris hop canker, pota<strong>to</strong> s<strong>to</strong>rage rot, root rot of<br />

many crops<br />

F. graminearum G. zeae numerous diseases of temperate and tropical cereals,<br />

e.g. pre- and post-emergence blights, root and<br />

foot rot, culm decay, head or kernel blight,<br />

ear scab and stalk rot.<br />

intercellular growth through the root tip in<strong>to</strong><br />

the xylem vessels (Bishop & Cooper, 1983). In<br />

planta, F. oxysporum exists mainly as microconidia<br />

which can spread in the xylem vessels. Sieve<br />

plates are overcome by germination and penetration,<br />

followed by the production of further<br />

microconidia. Wilting is caused by blockage of<br />

the xylem by fungal biomass and also by the<br />

accumulation of gums of plant origin, some of<br />

them released by fungal pectinase activity (Pegg,<br />

1985; Beckman, 1987). In cross section, infected<br />

stem bases show a browning of the vascular<br />

bundles. The host plant eventually dies because<br />

of water shortage (wilting), and after host death<br />

extensive colonization of host tissue and sporulation<br />

of F. oxysporum ensues. Extensive work<br />

is being carried out especially on molecular<br />

biological aspects of wilt diseases caused by


HYPOCREALES<br />

347<br />

Fig12.22 Fusarium culmorum. (a) Part of a sporodochium with clusters of phialides producing macroconidia. (b) Mature<br />

macroconidium from a two-week-old culture.The foot cell is indicated by an arrowhead. (c) Older macroconidia (eight-week-old)<br />

in which some cells have collapsed and the others have turned in<strong>to</strong> chlamydospores. (d) Chlamydospores formed by a hypha<br />

submerged in agar. (b d) <strong>to</strong> same scale.<br />

F. oxysporum, and this has been summarized by Di<br />

Pietro et al. (2003).<br />

The control of diseases caused by Fusarium is<br />

mainly through the use of resistant host plant<br />

cultivars (Beckman, 1987). The disastrous economic<br />

consequences of the outbreak and spread<br />

of Panama wilt of banana caused by F. oxysporum<br />

f. sp. cubense which attacked the popular banana<br />

variety ‘Gros Michel’ were overcome by switching<br />

production <strong>to</strong> clones of the resistant variety<br />

‘Cavendish’ which could be planted in<strong>to</strong> soils<br />

heavily contaminated with F. oxysporum f. sp.<br />

cubense (Moore et al., 2001). Another wellcharacterized<br />

example is <strong>to</strong>ma<strong>to</strong> wilt caused by<br />

F. oxysporum f. sp. lycopersici. Resistance seems <strong>to</strong><br />

be based mainly on major-gene resistance, and<br />

numerous resistance genes have been crossed<br />

in<strong>to</strong> the cultivated <strong>to</strong>ma<strong>to</strong> plant from wild<br />

relatives. These occur in several discrete gene<br />

clusters (Sela-Buurlage et al., 2001). A range of<br />

<strong>to</strong>ma<strong>to</strong> cultivars with resistance against one or<br />

more of the three races of F. oxysporum f. sp.<br />

lycopersici is available for planting.<br />

Another approach actively pursued at present<br />

is the biological control of F. oxysporum. An<br />

interesting strategy is the use of apathogenic<br />

strains of F. oxysporum which colonize the root<br />

cortex but not the vascular system, and therefore<br />

have an endophytic lifestyle. They inhibit infections<br />

by pathogenic strains. Various explanations<br />

have been offered, including competition for<br />

nutrients, the induction of systemic acquired<br />

resistance in plants pre-infected with the apathogenic<br />

strains, and direct competition between<br />

apathogenic and pathogenic strains in the root<br />

cortex (Larkin & Fravel, 1999; Bao & Lasarovits,<br />

2001). Other soil micro-organisms such as<br />

Pseudomonas spp. are also being developed as<br />

biocontrol agents.<br />

Myco<strong>to</strong>xins<br />

Some species of Fusarium (e.g. F. graminearum and<br />

F. moniliforme) are seed-borne, with mycelium<br />

being present inside the grain and often producing<br />

macroconidia on the surface. Infected seed<br />

not only ensures carry-over of the disease <strong>to</strong><br />

following seasons but can be a source of<br />

myco<strong>to</strong>xins if fed <strong>to</strong> lives<strong>to</strong>ck or consumed by<br />

humans. Myco<strong>to</strong>xins such as zearalenone, fumonisin,<br />

trichothecenes and vomi<strong>to</strong>xin (Fig. 12.23)


348 HYMENOASCOMYCETES: PYRENOMYCETES<br />

are produced by Fusarium spp. present on animal<br />

and human feeds<strong>to</strong>ck (Creppy, 2002; Moss, 2002).<br />

Zearalenone, an oestrogenic hormone from<br />

F. graminearum, causes vulvovaginitis and infertility<br />

in cattle and pigs, and trichothecenes such as<br />

T-2 <strong>to</strong>xin from several other Fusarium species<br />

cause <strong>to</strong>xic aleukia (reduction in white blood cell<br />

count) in farm animals and humans (Joffe, 1986;<br />

Moss, 2002).<br />

12.4.4 Cylindrocarpon (Nectriaceae)<br />

<strong>Fungi</strong> with Cylindrocarpon-type conidia are anamorphic<br />

states of species of Nectria sensu la<strong>to</strong>, now<br />

placed in a distinct genus Neonectria (Rossman<br />

et al., 1999; Mantiri et al., 2001). The macroconidia<br />

of Cylindrocarpon are hyaline, curved, transversely<br />

septate phialoconidia which resemble those of<br />

Fusarium but do not have the constricted basal<br />

foot cell characteristic of the latter (compare<br />

Figs. 12.21b and 12.20c). Some species also have<br />

microconidia and some have chlamydospores.<br />

About 120 taxa (species and varieties) have been<br />

described and about 50 have been linked <strong>to</strong><br />

Nectria sensu la<strong>to</strong>. Most species of Cylindrocarpon<br />

grow in soil and are saprotrophic or weakly<br />

parasitic. Cylindrocarpon destructans (formerly<br />

known as C. radicicola), whose teleomorph is<br />

Nectria radicicola (Samuels & Brayford, 1990),<br />

causes seedling blights and basal rots of bulbs,<br />

as well as root rots of various plants.<br />

Cylindrocarpon heteronema is the anamorph of<br />

Nectria galligena, the cause of apple and pear<br />

canker.<br />

12.5 Clavicipitales<br />

<strong>Fungi</strong> in this group have been and still are<br />

assigned <strong>to</strong> various groups (Spatafora &<br />

Blackwell, 1993; Rossman, 1996; Stensrud et al.,<br />

2005), notably a family (Clavicipitaceae) of the<br />

Hypocreales or a separate order (Clavicipitales).<br />

Whatever its taxonomic rank, this group<br />

contains fungi with several distinguishing<br />

characteristics. Perithecia develop on a fleshy<br />

stroma. The perithecial centrum contains a<br />

central basal mound from which the asci arise.<br />

Any paraphyses which develop are obliterated<br />

by crushing as the asci enlarge (White, 1997;<br />

Rossman et al., 1999). The asci have a well-defined<br />

Fig12.23 Common myco<strong>to</strong>xins produced by Fusarium spp. (a) The nonaketide zearalenone, a suspected carcinogen and<br />

oestrogen analogue produced by F. graminearum in maize and cattle feed. (b) Fumonisin B 1<br />

, a suspected cause of oesophageal<br />

cancer, produced by F. moniliforme infecting maize. (c) T-2 <strong>to</strong>xin, a highly cy<strong>to</strong><strong>to</strong>xic trichothecene (sesquiterpene derivative)<br />

produced by F. graminearum in various cereals. (d) Vomi<strong>to</strong>xin (desoxynivalenol), a trichothecene produced by F. culmorum and<br />

F. graminearum in maize and wheat.


CLAVICIPITALES<br />

349<br />

thick apical cap perforated by a narrow pore<br />

through which the ascospores are discharged,<br />

singly and successively. The ascospores are<br />

long, narrow and often multi-septate, breaking<br />

up in<strong>to</strong> part-spores. Most members (e.g. Claviceps,<br />

Balansia, Epichloe) are pathogens or endophytes<br />

of grasses, whereas Cordyceps parasitizes insects<br />

or fruit bodies of the hypogeous ascomycete<br />

Elaphomyces. Claviceps sclerotia are the source of<br />

<strong>to</strong>xic alkaloids, and grasses infected with endophytic<br />

Balansia may also be <strong>to</strong>xic <strong>to</strong> herbivorous<br />

insects and mammals. The grass host may thus<br />

be at least partially protected against insect<br />

herbivory, and the effects of alkaloids on<br />

grazing mammals can also be severe. The<br />

conidia of certain species of Cordyceps show<br />

promise as agents of biological control of insect<br />

pests (Evans, 2003). Some members of the<br />

Clavicipitales attack nema<strong>to</strong>des (Gams & Zare,<br />

2003). For example, Atricordyceps (now called<br />

Podocrella) is the teleomorph of Harposporium<br />

anguillulae (Samuels, 1983) (see Fig. 25.8), and<br />

phylogenetic analyses of nuclear ribosomal<br />

DNA have also placed the nema<strong>to</strong>phagous<br />

fungus Drechmeria coniospora (see Fig. 25.8) in<br />

the Clavicipitaceae (Gernandt & S<strong>to</strong>ne, 1999).<br />

The immunosuppressant drug cyclosporin A<br />

(Fig. 12.24a) is produced by Tolypocladium<br />

inflatum, anamorph of Cordyceps subsessilis<br />

(Hodge et al., 1996). A group of b-lactam antibiotics,<br />

the cephalosporins (Fig. 12.24b; see<br />

p. 302), are derived from the anamorphic<br />

Acremonium chrysogenum and A. salmosynnematum.<br />

Many other valuable secondary metabolites have<br />

been isolated from members of the Clavicipitales<br />

(Isaka et al., 2003).<br />

12.5.1 Claviceps<br />

There are over 40 species of Claviceps, all of which<br />

are parasitic on grasses, rushes and, occasionally,<br />

sedges (Alderman, 2003). The best known species<br />

is C. purpurea, the cause of ergot of grasses and<br />

cereals. Other economically important species<br />

are C. sorghi and C. africana which cause ergot<br />

of sorghum (Frederickson et al., 1991), C. paspali<br />

on Paspalum and C. fusiformis on pearl millet<br />

(Pennisetum typhoides). Claviceps purpurea occurs in<br />

temperate regions and has an exceptionally wide<br />

host range for a biotrophic pathogen, infecting<br />

over 400 grass species. The course of infection<br />

has been described by Luttrell (1980), Tenberge<br />

(1999) and Oeser et al. (2002).<br />

Life cycle<br />

The life cycle of C. purpurea is summarized in<br />

Fig. 12.25. The primary inoculum is an ascospore<br />

shot away from a perithecium which has developed<br />

from an overwintered sclerotium. The time<br />

of ascospore release coincides with anthesis in<br />

a susceptible host. Ascospores germinate on<br />

a grass stigma <strong>to</strong> form an intercellular mycelium<br />

which grows down <strong>to</strong> the base of the ovary<br />

<strong>to</strong>wards the vascular bundle of the floret stalk<br />

(rachilla), thus gaining access <strong>to</strong> the pho<strong>to</strong>synthetic<br />

products of the host. Subsequent growth is<br />

upwards and within a few days a conidial stroma<br />

develops beneath the ovary. A palisade of phialides<br />

lining labyrinthine chambers is formed<br />

from which a succession of unicellular, uninucleate<br />

conidia develops in a sugary syrup. This<br />

becomes visible on the grass florets as beads of<br />

liquid termed honeydew (Fig. 12.26b). The conidial<br />

stage was given the separate name Sphacelia<br />

segetum before its connection with ergot was<br />

unders<strong>to</strong>od. Honeydew contains glucose, fruc<strong>to</strong>se,<br />

sucrose and other sugars (Mower &<br />

Hancock, 1975a), and is attractive <strong>to</strong> insects,<br />

which feed on it and in so doing disperse conidia<br />

<strong>to</strong> healthy grass flowers, thus causing secondary<br />

infection. Infection of a grass flower by Claviceps<br />

results in increased translocation of water<br />

and sucrose <strong>to</strong>wards the diseased flower, and<br />

infected flowers are more effective at acquiring<br />

pho<strong>to</strong>synthetic products from the host than<br />

uninfected flowers (Parbery, 1996b). Within the<br />

infected host tissue, conversion of host-derived<br />

sucrose <strong>to</strong> mono-, di- and oligo-saccharides by the<br />

fungus creates a continuing sink for sucrose<br />

translocation, and evaporation at the surface of<br />

the diseased grain results in increased osmotic<br />

concentration of the sugars, possibly accelerating<br />

the rate of translocation (Mower & Hancock,<br />

1975b). The high osmotic concentration prevents<br />

conidial germination until the honeydew has<br />

been diluted.<br />

As infection proceeds, the entire ovary is<br />

pushed upwards by the developing fungal tissue


350 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.24 Important metabolites from<br />

Clavicipitales. (a) Cyclosporin A, produced by<br />

Tolypocladium inflatum.This is used extensively as<br />

an immunosuppressive drug, e.g. after organ<br />

transplantations. (b) Cephalosporin C,<br />

a b-lactam antibiotic with activity against<br />

Gram-positive and Gram-negative bacteria.<br />

and sits like a cap over it. The ovary, which would<br />

normally develop in<strong>to</strong> a caryopsis filled with<br />

grain, becomes replaced by fungal tissue. For this<br />

reason the disease caused by C. purpurea has been<br />

termed a replacement disease (Luttrell, 1980).<br />

The fungal structure which develops is considerably<br />

longer than the ovary which it replaces and<br />

it differentiates in<strong>to</strong> a sclerotium up <strong>to</strong> 3 cm in<br />

length, the foot of which continues <strong>to</strong> obtain<br />

nutrients via the vascular connection in the host<br />

rachilla. The sclerotium is made up of three<br />

distinct layers: a thin purplish-brown rind, a<br />

discontinuous layer of mealy white tissue, and<br />

a central layer of translucent gelatinous tissue<br />

(Luttrell, 1980).<br />

Ergot is the French name for a cock’s spur,<br />

referring <strong>to</strong> the curved, banana-shaped sclerotia<br />

which project from the inflorescence of infected<br />

grasses and cereals in late summer (Fig. 12.26a).<br />

The sclerotia fall <strong>to</strong> the ground and overwinter<br />

near the surface of the soil. They need a period of<br />

low temperature before they can develop further.<br />

A chilling period of 0°C for at least 25 days<br />

is optimal for further development. The main<br />

reserve substance of the sclerotium is lipid,<br />

which may account for 50% of the dry weight.<br />

It is likely that the chilling period is necessary<br />

before enzymes capable of mobilizing the lipid<br />

reserves develop (Cooke & Mitchell, 1970).<br />

Sclerotia do not remain viable in soil in the<br />

field for more than a few months, often being<br />

invaded by fungi, bacteria, mites and insects<br />

(Cunfer & Seckinger, 1977). The following<br />

summer, sclerotia develop one or more perithecial<br />

stromata (clavae) about 1 2 cm high, shaped<br />

like miniature drumsticks (Fig. 12.26c). The<br />

perithecial stromata are positively pho<strong>to</strong>tropic<br />

(Hadley, 1968).<br />

The enlarged spherical head or capitulum<br />

contains a number of perithecia which are embedded<br />

in the stroma, each surrounded by a<br />

distinct perithecial wall (Fig. 12.27a). The cy<strong>to</strong>logical<br />

details of perithecial development have<br />

been studied in C. purpurea by Killian (1919) and<br />

in C. microcephala (regarded by some as a form of<br />

C. purpurea) by Kulkarni (1963). In the outer layers<br />

of the head of the perithecial stroma, clubshaped<br />

multinucleate antheridia and ascogonia<br />

undergo plasmogamy. Ascogenous hyphae made<br />

up of predominantly binucleate segments<br />

develop from the base of the ascogonium, and<br />

the tips of the ascogenous hyphae form croziers


CLAVICIPITALES<br />

351<br />

Fig12.25 The homothallic life cycle of<br />

Claviceps purpurea growing on rye (Secale<br />

cereale). A sclerotium fallen <strong>to</strong> the ground<br />

will produce a perithecial stroma after<br />

overwintering. In the perithecia,<br />

plasmogamy (P), karyogamy (K) and<br />

meiosis (M) give rise <strong>to</strong> filamen<strong>to</strong>us<br />

ascospores, each with numerous haploid<br />

nuclei (small open circles). Ascospores<br />

germinate on the stigmata of the rye<br />

ovary, and mycelium penetrates <strong>to</strong> the<br />

ovary stalk (rachilla). Infections develop<br />

both as conidial regions and, lower down,<br />

as a sclerotium initial.The developing<br />

sclerotium pushes the ovary upwards,<br />

andthisremainsasacapabovethe<br />

conidial stroma.The stroma contains<br />

cavities lined by phialides which produce<br />

phialoconidia of the Sphacelia segetum<br />

type.These accumulate in beads of a<br />

sugary liquid (honeydew).Conidia are<br />

carried by insects <strong>to</strong> fresh rye stigmata<br />

and initiate secondary infections. Later in<br />

the season, the sclerotia enlarge and<br />

become visible as ergots. Some images<br />

redrawn from Luttrell (1980).<br />

with binucleate penultimate segments. The<br />

penultimate cell elongates <strong>to</strong> form the ascus,<br />

and fusion between the two nuclei occurs. There<br />

are numerous asci in each perithecium, each<br />

containing a bundle of eight filiform ascospores.<br />

The ascus bears a conspicuous perforated cap at<br />

its tip (Fig. 12.27c).<br />

Successful infection of rye (Secale cereale) from<br />

cultures derived from a single ascospore show<br />

that C. purpurea is homothallic. Despite this,<br />

genetic recombination is possible through<br />

heterokaryosis and parasexual reproduction<br />

(Tudzynski, 1999). Curiously, sclerotia are<br />

frequently formed from heterokaryotic mycelia,<br />

indicating multiple infection of the grass flower.<br />

Although common on rye and some other<br />

cereals in Europe and North America, C. purpurea<br />

is not usually troublesome on cereals in Britain.<br />

In the occasional years in which its incidence is<br />

high, there is a correlation with high relative<br />

humidity and low maximum temperature in<br />

June, which probably prolongs the period during


352 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.26 Claviceps purpurea. (a) Head of rye (Secale cereale) bearing several sclerotia (ergots) of Claviceps purpurea. (b) Rye<br />

inflorescence at anthesis bearing two drops (arrowed) of the honeydew or Sphacelia conidial state of C. purpurea. A fly has landed<br />

near these drops. (c) Germinated sclerotium showing several stalked perithecial stromata.<br />

which the host grass flowers are open and<br />

therefore susceptible <strong>to</strong> infection.<br />

Ergotism<br />

The effect of infection on the host can result<br />

in yield reduction by as much as 80% of seeds.<br />

Severe though this reduction in crop yield<br />

is, the consequences of consumption of ergotcontaminated<br />

grain can be disastrous <strong>to</strong> herbivorous<br />

animals and humans alike. The purple<br />

sclerotia contain a number of <strong>to</strong>xic alkaloids<br />

(Buchta & Cvak, 1999) and if they are eaten they<br />

can cause severe illness and sometimes death.<br />

Even at relatively low concentration there may<br />

be severe effects on feed refusal, lack of weight<br />

gain in farm animals and on reduced fertility,<br />

resulting in part from agalactia, the inability <strong>to</strong><br />

produce sufficient milk <strong>to</strong> nourish the young<br />

(Shelby, 1999). One effect of the <strong>to</strong>xins is <strong>to</strong><br />

constrict the blood vessels, and the impaired<br />

circulation may result in gangrene or loss of<br />

limbs. Gruesome descriptions of the symp<strong>to</strong>ms<br />

on humans have been related by several contemporary<br />

authors, e.g. Sidney (1846) who wrote:<br />

The medical effects of ergot, in small doses, have<br />

already been noticed as being extremely powerful,<br />

but if taken <strong>to</strong> any extent its results on the animal<br />

frame are truly awful. This has been proved by<br />

numerous experiments, of which Professor Henslow<br />

gives a most striking account in his most valuable<br />

notice of this disease; <strong>to</strong> which he adds a proper<br />

caution against their repetition now the question is<br />

settled. Animals which refused ergot mixed with<br />

their food have been compelled <strong>to</strong> swallow it, and it<br />

reduced them <strong>to</strong> a wretched condition. It was tried<br />

upon pigs, and also upon poultry, and the<br />

consequences were sickness, gangrene, and<br />

inflamma<strong>to</strong>ry action so intense, that the flesh<br />

actually sloughed away. In some cases, the limbs<br />

rotted off, and no description of animal suffering<br />

has ever exceeded the direful ills thus inflicted.<br />

These experiments were made with a view <strong>to</strong><br />

determine whether the ergot of rye, constantly<br />

ground up with the flour in some parts of France,<br />

might not be the cause of the gangrenous disease<br />

so prevalent amongst the poor in certain districts.<br />

The symp<strong>to</strong>ms of these epidemic diseases are


CLAVICIPITALES<br />

353<br />

and dumb, and, besides, lost a limb which actually<br />

rotted off, precisely in the same way as the limbs of<br />

the animals which were compelled <strong>to</strong> swallow the<br />

experimental ergot.<br />

Another effect is on the nervous system,<br />

resulting in convulsions, hallucinations and<br />

burning sensations. In the Middle Ages the<br />

symp<strong>to</strong>ms of ergotism were called ‘St<br />

Anthony’s fire’ and there are numerous records<br />

of out-breaks of the disease (see Ramsbot<strong>to</strong>m,<br />

1953; Fuller, 1969; Ma<strong>to</strong>ssian, 1989). Ma<strong>to</strong>ssian<br />

(1989) has outlined some of the social consequences<br />

of ergotism, e.g. in depressing population<br />

growth after the plague outbreaks in<br />

Europe, and in provoking witch trials in North<br />

America when women were accused of and<br />

executed for bewitching people who were probably<br />

suffering from ergot-induced food poisoning<br />

and hallucinations.<br />

With improved grain-cleaning techniques<br />

and a switch in carbohydrate consumption<br />

from rye <strong>to</strong> wheat, maize and pota<strong>to</strong>es, the<br />

disease is now rare in humans. Cattle and sheep<br />

which have eaten sclerotia from pasture grasses,<br />

or pigs and horses fed on ergot-contaminated<br />

grain, are still affected and if pregnant animals<br />

are involved there is a risk of abortion. However,<br />

these problems are now relatively rare.<br />

Fig12.27 Claviceps purpurea. (a) L.S. perithecial stroma.<br />

(b) T.S. young sclerotium showing the formation of<br />

phialoconidia on the surface. (c) Ascus and ascospores.<br />

Note the cap of the ascus.<br />

dreadful, and there seems <strong>to</strong> be very little doubt that<br />

the suspicions as <strong>to</strong> their originating from ergotted<br />

flour of rye are correct. Tessier, who has paid great<br />

attention <strong>to</strong> the subject, mentions a case which<br />

came under his own observation. A family were in<br />

a state of great destitution, and the father begged of<br />

a neighbouring farmer a quantity of ergotted rye <strong>to</strong><br />

supply the urgent calls of his distressed family for<br />

food. The farmer gave it him, but added that he was<br />

afraid it was not wholesome. Still the calls of hunger<br />

prevailed, and in the face of this caution it was<br />

eaten. The result was the death of the father,<br />

mother, and five of the children out of seven. Two<br />

survived, but one of them became subsequently deaf<br />

Alkaloids of Claviceps purpurea<br />

Ergot alkaloids are used in human medicine. ‘No<br />

other class of compounds exhibits such a wide<br />

spectrum of structural diversity, biological activity<br />

and therapeutic uses as ergot derivatives’<br />

(Křen & Cvak, 1999). This is because the tetracyclic<br />

ergoline ring structure (Fig. 12.28) mimics<br />

several neurotransmitter molecules such as<br />

noradrenaline, dopamine and sero<strong>to</strong>nin<br />

(Mantegani et al., 1999). Two alkaloids, ergometrine<br />

and ergotamine, are of special importance<br />

and are produced by field strains of C. purpurea<br />

(Pažou<strong>to</strong>vá et al., 2000). Ergometrine causes<br />

constriction of smooth muscle tissues. The<br />

name ergometrine is derived from endometrium,<br />

the lining of the uterus, because the<br />

drug is used <strong>to</strong> stimulate uterine contraction.<br />

Ergotamine can similarly accelerate uterine<br />

contraction and is used as a vasoconstric<strong>to</strong>r


354 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.28 Biosynthesis of alkaloids in ergots of Claviceps purpurea. Starting from tryp<strong>to</strong>phan and the mevalonic acid-derived metabolite<br />

dimethylallylpyrophosphate, dimethylallyltryp<strong>to</strong>phan is synthesized. Ring closure, decarboxylation and addition of a methyl<br />

group from methionine (arrowhead) gives the simplest alkaloids, agroclavine and D-lysergic acid.The latter is then derivatized at its<br />

C-8 position <strong>to</strong> give more complex alkaloids such as ergometrine and ergotamine. LSD (D-lysergic acid diethylamide) is a semisynthetic<br />

derivative of D-lysergic acid.<br />

and a haemostatic drug. It was also the first<br />

medication available against migraine (Eadie,<br />

2004). A third ergot-derived drug is lysergic acid<br />

diethylamide (LSD), a synthetic derivative of<br />

lysergic acid. As is well known, this drug has<br />

hallucinogenic properties and is associated with<br />

euphoria (Minghetti & Crespi-Perellino, 1999).<br />

Numerous chemical derivatives have been<br />

synthesized, and ergot alkaloids are lead structures<br />

for drugs against a range of illnesses,<br />

including Parkinson’s disease and Alzheimer’s<br />

disease.<br />

The biosynthesis of ergot alkaloids in<br />

C. purpurea from tryp<strong>to</strong>phan, mevalonic acid<br />

and methyl groups donated by methionine<br />

leads <strong>to</strong> the ergoline ring structure (Fig. 12.28).<br />

The simplest alkaloids are the clavine alkaloids<br />

such as agroclavine, but most of the pharmaceutically<br />

important ones are based on D-lysergic<br />

acid because it is the carboxylic acid group at C-8<br />

in D-lysergic acid which is further substituted.<br />

Simple substituted D-lysergic acid derivatives<br />

are ergometrine and the semi-synthetic LSD<br />

(Fig. 12.28). In the most complex ergot alkaloids,<br />

the ergopeptines, three amino acids are added<br />

<strong>to</strong> D-lysergic acid by a non-ribosomal peptide<br />

synthase, followed by ring closure. Ergotamine<br />

is a commonly found example of this type of<br />

alkaloid in C. purpurea (Fig. 12.28). Reviews of<br />

alkaloid biosynthesis have been written by<br />

Flieger et al. (1997) and Tudzynski et al. (2001).<br />

A concise general treatment is that by Lohmeyer<br />

and Tudzynski (1997).<br />

The ergot of commerce is produced by cultivating<br />

the fungus on rye, and profitable crops<br />

of ergot sclerotia are obtained, often on specially<br />

bred strains of male-sterile rye in Eastern Europe,<br />

Spain and Portugal (Németh, 1999). Rye flowers<br />

are infected with a suspension of conidia from<br />

strains of C. purpurea selected for high yield of<br />

certain alkaloids. Alkaloids are extracted from<br />

harvested sclerotia. Alkaloids are also extracted<br />

from special strains of C. purpurea grown saprotrophically<br />

in deep fermentation. About 60% of


CLAVICIPITALES<br />

355<br />

alkaloid production is currently derived from<br />

fermentations, the rest from harvested sclerotia.<br />

The advantages of saprotrophic fermentation<br />

are that the process can be closely controlled<br />

and the alkaloids produced are less variable than<br />

those from harvested ergots. A disadvantage is<br />

that the ability <strong>to</strong> produce alkaloids in economically<br />

significant amounts is variable and may be<br />

lost on prolonged cultivation. Nevertheless,<br />

the market share of alkaloids produced by<br />

fermentation is currently increasing (Tudzynski<br />

et al., 2001).<br />

Control of Claviceps<br />

The control of ergot in cereals is difficult.<br />

Although several techniques are available, none<br />

is completely effective. Use of ergot-free seed<br />

would reduce infection, but inoculum may<br />

survive from a previous crop. It can also be<br />

provided by wild grasses bordering the field<br />

because C. purpurea strains have wide host<br />

ranges. Deep ploughing, which buries the sclerotia,<br />

and crop rotation involving a non-cereal<br />

are also helpful. Systemic fungicides would need<br />

<strong>to</strong> be applied in sufficient amounts <strong>to</strong> produce<br />

an effective concentration at the surface of the<br />

ovary, and they have been used <strong>to</strong> control ergot<br />

in seed crops of Kentucky bluegrass, Poa pratensis<br />

(Schulz et al., 1993). <strong>Fungi</strong>cide sprays are used at<br />

present <strong>to</strong> control C. africana on sorghum in<br />

Australia (Ryley et al., 2003) but not against<br />

C. purpurea on cereals.<br />

Other Claviceps spp.<br />

Some other species of Claviceps differ in significant<br />

ways from C. purpurea. For example,<br />

C. fusiformis has two synanamorphs, a macroand<br />

a micro-conidial state. Claviceps africana and<br />

C. paspali may produce secondary phialoconidia<br />

and these may develop in sufficient quantity on<br />

the surface of the conidial stroma <strong>to</strong> be capable<br />

of dispersal by wind (Luttrell, 1977; Frederickson<br />

& Mantle, 1989; Alderman, 2003). Claviceps paspali<br />

which infects dallisgrass (Paspalum dilatatum) is<br />

the source of alkaloids such as paspalic acid and<br />

its derivatives. Ingestion of its sclerotia causes<br />

paspalum staggers in sheep. The salt marsh grass<br />

Spartina anglica often shows heavy infection by<br />

a specialized variety, C. purpurea var. spartinae<br />

(Plate 5e). This fungus appears <strong>to</strong> be adapted <strong>to</strong><br />

an aquatic environment. Its unusually slender<br />

sclerotia float on the surface of sea water whilst<br />

those of other forms of C. purpurea sink. The high<br />

levels of infection may be related <strong>to</strong> the fact that<br />

S. anglica, an allopolyploid grass of recent origin,<br />

is genetically uniform. Despite the heavy infection,<br />

seed production by the host plant is not<br />

severely affected (Raybould et al., 1998; Duncan<br />

et al., 2002). In contrast, C. phalaridis, which is<br />

endemic in Australia on the introduced pasture<br />

grass Phalaris tuberosa, is systemic and when its<br />

mycelium penetrates the inflorescence, sclerotia<br />

are formed in all the florets, rendering the host<br />

plant sterile (Walker, 2004). Its systemic habit<br />

is shared by other clavicipitaceous endophytes<br />

such as Epichloe (see below).<br />

12.5.2 Epichloe<br />

There are about 10 biological species (i.e. mating<br />

populations) of Epichloe (Gr. epi ¼ on, upon;<br />

chloë ¼ young shoots of grass) mainly infecting<br />

cool-season grasses with the C 3 pho<strong>to</strong>synthetic<br />

pathway (Leuchtmann, 2003). They grow in<br />

nature as biotrophic, systemic, parasitic or<br />

symbiotic endophytes in grass shoots, forming<br />

at first conidial, then perithecial stromata<br />

around the uppermost leaf sheaths of tillers<br />

containing floral primordia. Anamorphic relatives,<br />

now classified as species of<br />

Neotyphodium (previously Acremonium Section<br />

Albo-lanosum; Glenn et al., 1996), are symp<strong>to</strong>mless<br />

endophytes. Neotyphodium-infected grasses<br />

contain ergot alkaloids and other myco<strong>to</strong>xins<br />

which are injurious <strong>to</strong> herbivorous insects and<br />

mammals and cause economic damage (Schardl,<br />

1996; Kuldau & Bacon, 2001; Clay & Schardl,<br />

2002).<br />

Epichloe typhina (sensu la<strong>to</strong>) causes ‘choke’ of<br />

pasture grasses and is common on grasses such<br />

as Dactylis, Holcus and Agrostis. However, forms<br />

of Epichloe on certain hosts are distinct from<br />

E. typhina in dimensions of stromata, ascospores,<br />

ascospore septation, and in molecular characteristics.<br />

They have now been accorded different<br />

species names. The specific name typhina should<br />

be applied <strong>to</strong> the forms on eight genera of<br />

grasses including Dactylis, whilst the name


356 HYMENOASCOMYCETES: PYRENOMYCETES<br />

E. baconii has been given <strong>to</strong> the fungus on Agrostis<br />

s<strong>to</strong>lonifera, and E. clarkii <strong>to</strong> that on Holcus lanatus<br />

(White, 1993). The form on Festuca rubra and F.<br />

valesiaca is E. festucae (Leuchtmann et al., 1994).<br />

The uppermost leaf sheath of flowering tillers<br />

becomes surrounded by a white mass of mycelium<br />

2 cm or more in length, and at the surface<br />

small unicellular phialoconidia are produced<br />

(Fig. 12.29b). These conidia function as spermatia<br />

(see below). Later, the conidial stroma becomes<br />

thicker and turns orange in colour as perithecia<br />

are formed (Plate 5f). The perithecia produce<br />

numerous asci, each with a well-defined apical<br />

cap, and containing eight long narrow ascospores<br />

which may break up within the ascus <strong>to</strong><br />

form part-spores (Fig. 12.29d). The mycelium is<br />

for the most part intercellular, unbranched and<br />

mainly located in the pith, although intracellular<br />

penetration of the vascular bundles is found<br />

in the region of the inflorescence primordium.<br />

Perithecial stromata are formed only on tillers<br />

containing inflorescence primordia, and by<br />

manipulating incubation conditions it has been<br />

shown that the formation of stromata is correlated<br />

directly with the presence of an inflorescence<br />

primordium rather than with external<br />

conditions (Kirby, 1961).<br />

Epichloe typhina is heterothallic with a unifac<strong>to</strong>rial<br />

(bipolar) mating system (White & Bultman,<br />

1987). Throughout its range Epichloe is attacked<br />

by a parasitic fly, Botanophila phrenione (Phorbia<br />

phrenione), which feeds on conidia, conidiophores<br />

and hyphae. The relationship is a symbiotic one.<br />

Before laying eggs, female flies feed on conidia<br />

and hyphae from conidial stromata. Possibly they<br />

are attracted by the white colour and distinctive<br />

smell of the stroma (Leuchtmann, 2003). The<br />

conidia remain viable after passing through the<br />

gut of the flies. After laying an egg in a fresh<br />

conidial stroma an ovipositing female shows an<br />

unusual but characteristic pattern of behaviour,<br />

walking in a linear or spiral path around the<br />

conidial stroma whilst dragging its abdomen and<br />

depositing a trail of faecal material with viable<br />

conidia on the receptive hyphae of its surface,<br />

so spermatizing them. This track is later marked<br />

by the development of perithecia (Bultman<br />

et al., 1998).<br />

Perithecial on<strong>to</strong>geny has been studied by<br />

White (1997). The perithecial primordium develops<br />

as a cavity lined by inwardly directed<br />

branched hyphae. In E. typhina, at the base of the<br />

cavity a mound of ascogenous tissue appears.<br />

This is made up of ascogenous hyphae with<br />

croziers and with lateral paraphyses, but the<br />

paraphyses do not persist as the asci mature. The<br />

perithecial ostioles protrude above the surface of<br />

the stroma and are lined by curved periphyses<br />

(Fig. 12.29c). The apical apparatus of the ascus<br />

consists of a thickened ring pierced by a narrow<br />

canal continuous with the cy<strong>to</strong>plasm of the ascus<br />

(Figs. 12.29d, 12.30a) and through the canal the<br />

ascospores are discharged singly, one after<br />

another.<br />

Ascospores may segment in<strong>to</strong> part-spores<br />

within the ascus or after discharge. It has been<br />

claimed that they never germinate directly (i.e.<br />

by germ tube), but only by the production of<br />

conidia from narrow tapering phialides (Figs.<br />

12.30b,c; Bacon & Hin<strong>to</strong>n, 1988). However, our<br />

own observations on ascospores of E. typhina<br />

from D. glomerata show that direct germination<br />

may occur (Fig. 12.30d). Primary conidia may<br />

germinate directly or by forming secondary<br />

conidia, a process described as microcyclic conidiation<br />

(Bacon & Hin<strong>to</strong>n, 1991). Tertiary conidia<br />

may also develop from secondary conidia.<br />

Attempts <strong>to</strong> infect grasses from ascospores<br />

have generally been unsuccessful, so it is likely<br />

that infection is by conidia. After allowing<br />

ascospores <strong>to</strong> be discharged close <strong>to</strong> emerging<br />

inflorescences of uninfected Lolium perenne<br />

plants, about 12% of the seeds gave rise <strong>to</strong><br />

infected progeny, but whether infection was<br />

directly from ascospores or from primary or<br />

secondary conidia was not determined (Chung &<br />

Schardl, 1997a). Experimentally, it has not been<br />

possible <strong>to</strong> infect developing seeds of Dactylis,<br />

and the only effective method of infection is by<br />

application of ascospores or conidia <strong>to</strong> the cut<br />

ends of green stubble (Western & Cavett, 1959).<br />

If this is the natural route of infection in other<br />

grasses, it may explain the greater incidence of<br />

the disease in Agrostis in heavily grazed pastures<br />

(Bradshaw, 1959).<br />

The time of release of ascospores coincides<br />

with the emergence of larvae of the parasitic fly


CLAVICIPITALES<br />

357<br />

Fig12.29 Epichloe baconii. (a) T.S. stem and leaf<br />

sheath of Agrostis surrounded by a perithecial stroma.<br />

Note the axillary shoots between the leaf sheath and<br />

the stem. (b) Part of the conidial stroma. (c) A single<br />

perithecium. Note the periphyses lining the ostiole.<br />

(d) Ascus and ascospores. Note the apical apparatus<br />

of the ascus.<br />

which may feed on perithecia and so reduce<br />

ascospore production (Welch & Bultman, 1993).<br />

12.5.3 Epichloe-related grass endophytes<br />

White (1988) has classified the relationships<br />

between Epichloe or related endophytes and<br />

their hosts in<strong>to</strong> three types. In type I associations,<br />

perithecial stromata are formed on the inflorescences<br />

of most if not all infected individuals<br />

so that flowering of the host is suppressed.<br />

Dactylis glomerata and Agrostis tenuis harbour this<br />

type of association which should be regarded<br />

as parasitic. In type II associations, stromata<br />

are formed on only a few (1 10%) of infected<br />

individuals in a population, although 50 75% of<br />

the population may contain the infection. This<br />

type of association has been found only in the<br />

sub-family Festucoideae. Agrostis hiemalis, Bromus<br />

anomalus and Elymus canadensis have associations<br />

of this type. The endophyte is probably spread<br />

by clonal (i.e. vegetative) growth of the infected<br />

host, by ascospores (contagious or horizontal<br />

transmission) and by seed transmission (vertical<br />

transmission). In type III associations, stromata<br />

are not formed on infected plants and apparently<br />

are never produced. Such associations<br />

have been found only in festucoid grasses including<br />

tall fescue (Festuca arundinacea) and perennial<br />

ryegrass (Lolium perenne). In many of these grasses<br />

over 90% of individuals are infected. In this type<br />

of association the endophytes rely on vertical<br />

transmission which involves mycelial growth<br />

from the parent plant <strong>to</strong> the embryo within the<br />

seed (White et al., 1991). Since the host plant is


358 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.30 Epichloe typhina. (a) Tip of ascus showing the thickened apical ring pierced by a cy<strong>to</strong>plasmic canal. (b) Discharged<br />

ascospore which has germinated <strong>to</strong> produce several phialides. (c) Phialides with primary conidia. (d) Ascospore showing<br />

direct germination by hyphal growth. Some phialides are also visible.<br />

not detrimentally affected and indeed may<br />

benefit by experiencing reduced herbivory, the<br />

relationship can be regarded as mutualistic.<br />

The anamorphic states of Epichloe spp., now<br />

classified as species of Neotyphodium, grow as<br />

endophytes, i.e. symp<strong>to</strong>mless symbionts, of<br />

many grasses. Most species of Neotyphodium<br />

produce phialoconidia in labora<strong>to</strong>ry culture<br />

and N. typhinum, the anamorph of E. typhina,<br />

has been reported <strong>to</strong> form conidiophores on the<br />

phylloplane (leaf surfaces) of Poa rigidifolia and<br />

Agrostis hiemalis, although it is unclear whether<br />

or not they are involved in horizontal transmission<br />

in nature (White et al., 1996). What is certain<br />

is that these endophytes are seed-borne and<br />

are vertically transmitted. The leaf sheaths of<br />

infected plants show a characteristic fine, intercellular,<br />

infrequently branched, con<strong>to</strong>rted mycelium<br />

with lipid contents, running parallel <strong>to</strong> the<br />

vascular bundles, and following the longitudinal<br />

cell walls of the inner epidermis of the lower<br />

leaf sheaths (Fig. 12.31). Endophyte mycelium<br />

is less extensive in leaf blades and it is believed<br />

that the ligule, lacking intercellular spaces,<br />

may be a barrier <strong>to</strong> spread (Hin<strong>to</strong>n & Bacon,<br />

1985; Christensen et al., 2002). The mycelium of<br />

Neotyphodium appears indistinguishable from<br />

that of Epichloe. At flowering, the Neotyphodium<br />

mycelium grows upwards through the stem,<br />

extending in<strong>to</strong> the inflorescence and infecting<br />

the embryos of the developing seeds so that<br />

a high proportion of them are infected (White<br />

et al., 1991). In natural infections of Lolium perenne<br />

by N. lolii, the mycelium is scanty in vascular<br />

tissues but is occasionally found in smaller<br />

vascular bundles (Christensen et al., 2001).<br />

Alkaloids and endophytism<br />

Intense interest in Neotyphodium has developed<br />

since the discovery that consumption of<br />

endophyte-infected grass is associated with disorders<br />

of grazing lives<strong>to</strong>ck caused by myco<strong>to</strong>xins,<br />

including ergot alkaloids. Two associations<br />

(symbiota) have been particularly well investigated.<br />

Ingestion of tall fescue (Festuca arundinacea)<br />

infected with N. coenophialum causes fescue<br />

<strong>to</strong>xicosis in cattle and horses in the Southeastern<br />

USA (Bacon et al., 1977; Blodgett, 2001). The


CLAVICIPITALES<br />

359<br />

Fig12.31 Endophytic hyphae of<br />

Neotyphodium lolii growing<br />

intercellularly between epidermal<br />

cells of Lolium perenne.<br />

Pho<strong>to</strong>graphed from material kindly<br />

provided by P. J.Fisher.<br />

economic impact of fescue <strong>to</strong>xicosis <strong>to</strong> lives<strong>to</strong>ck<br />

producers in the USA has been estimated at<br />

$50 200 million annually (Siegel et al., 1984).<br />

Perennial ryegrass (Lolium perenne) containing<br />

N. lolii is associated with ryegrass staggers in<br />

sheep in New Zealand (Fletcher & Harvey, 1981).<br />

Various alkaloids are present in Neotyphodiuminfected<br />

grasses. These belong <strong>to</strong> several groups,<br />

including the ergopeptide-type ergot alkaloids<br />

which we have already encountered in Claviceps<br />

purpurea (Figs. 12.28, 12.32a), the lolitrems<br />

(Fig. 12.32b), the lolines (Fig. 12.32c) and peramine<br />

(Fig. 12.32d). The primary causal agent<br />

of ryegrass staggers is the neuro<strong>to</strong>xin lolitrem B<br />

(Siegel & Bush, 1997; Kuldau & Bacon, 2001).<br />

In addition <strong>to</strong> their <strong>to</strong>xicity <strong>to</strong> mammals,<br />

endophyte-infected grasses have deleterious<br />

effects on insects feeding on them. The loline<br />

alkaloids are primarily insecticidal. The endophytes<br />

are therefore regarded as mutualistic<br />

symbionts with their grass hosts, protecting<br />

them against mammalian and insect herbivory<br />

whilst themselves gaining nutrients and a means<br />

of dissemination (Clay, 1988; Schardl & Clay,<br />

1997). Endophyte-infected grasses are also more<br />

resistant than uninfected hosts against attack by<br />

certain fungal pathogens (Christensen, 1996) and<br />

nema<strong>to</strong>des.<br />

There are other effects of infection on the<br />

grass hosts. Infection causes enhanced production<br />

of biomass, increased tillering, drought<br />

resistance and competitiveness. These are valuable<br />

attributes of turf grasses and attempts have<br />

been made <strong>to</strong> enhance the agronomic value of<br />

turf grasses by deliberately infecting them with<br />

endophytic fungi (Bacon et al., 1997), including<br />

genetically modified strains. Attempts have also<br />

been made <strong>to</strong> free seeds and seedlings of pasture<br />

grasses from endophytes by fungicidal or heat<br />

treatment. Prolonged s<strong>to</strong>rage may also achieve<br />

this goal because the endophytes do not retain<br />

viability for as long as the seeds. However,<br />

endophyte-free plants are often less competitive<br />

than their infected counterparts.<br />

The origin of grass endophytes<br />

The most convincing evidence that Neotyphodium<br />

species are the anamorphic state of Epichloe is<br />

molecular (Glenn et al., 1996; Kuldau et al., 1997).<br />

Some of the endophytic Neotyphodium species are<br />

believed <strong>to</strong> be directly related <strong>to</strong> a species of<br />

Epichloe, e.g. N. lolii which is apparently derived<br />

from E. festucae, whilst others are of hybrid<br />

origin (Tsai et al., 1994; Schardl & Moon, 2003).<br />

Interspecific heterokaryon formation has been<br />

observed in culture (Chung & Schardl, 1997b).<br />

Hyphal anas<strong>to</strong>mosis and heterokaryosis may<br />

take place within grass shoots infected with<br />

more than one species of Epichloe, and the parasexual<br />

origin of some Neotyphodium species is<br />

a possibility (Tredway et al., 1999).


360 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.32 Alkaloids produced by Neotyphodium endophytes in grasses. (a) Ergovaline, an ergopeptide similar <strong>to</strong> ergotamine<br />

(see Fig.12.28). (b) Lolitrem B, an alkaloid derived from geranylgeranylpyrophosphate and tryp<strong>to</strong>phan.This substance is neuro<strong>to</strong>xic<br />

<strong>to</strong> grazing animals, causing ryegrass staggers. (c) Loline, an insecticidal pyrrolizidine alkaloid derived from ornithine and<br />

S-adenosylmethionine. (d) Peramine, an insecticidal pyrrolopyrazine-type alkaloid.<br />

Other evidence comes from the distribution<br />

of alkaloids in species of Epichloe and<br />

Neotyphodium. Epichloe festucae is the only sexual<br />

species known <strong>to</strong> synthesize representatives of<br />

three classes of alkaloids, namely ergovaline,<br />

lolitrem B and lolines (Fig. 12.32). Neotyphodium<br />

lolii produces the same three alkaloids and it<br />

seems reasonable <strong>to</strong> assume that E. festucae<br />

was the ances<strong>to</strong>r that contributed the genes for<br />

synthesis of these three alkaloids (Clay & Schardl,<br />

2002).<br />

12.5.4 Cordyceps and its anamorphs<br />

There are about 400 500 species of Cordyceps<br />

(Kobayasi, 1982; Liu et al., 2002) with an epicentre<br />

of species diversity in northeastern Asia and<br />

Japan. There are 29 species in the United States<br />

and Canada, and 18 species in Europe (Humber,<br />

2000). Most species are necrotrophic parasites<br />

of insect adults, larvae or pupae, but several<br />

species grow on the ascocarps of Elaphomyces<br />

(Mains, 1957). There is a wide range of insect<br />

hosts including moths, ants, beetles, and cicadas.<br />

Spiders are also attacked. Within the dead body<br />

of an infected insect a mass of mycelium or a<br />

sclerotium develops, and from this a perithecial<br />

stroma grows out. The perithecial stroma is<br />

usually fleshy and brightly coloured. It may<br />

bear perithecia over its entire surface or they<br />

may be restricted <strong>to</strong> an upper or lateral portion<br />

so that there is a stalk region lacking perithecia.<br />

The perithecia are embedded in stromatal tissue<br />

and tightly packed <strong>to</strong>gether. They are elongate,<br />

with protruding ostioles. They contain numerous<br />

narrowly cylindrical asci, each with a conspicuous<br />

swollen cap pierced by a narrow canal.<br />

There are 4 8 long cylindrical ascospores which<br />

are typically divided by transverse septa in<strong>to</strong><br />

cylindrical or fusoid part-spores which may<br />

number 16, 32, 64 or 128 (Hywel-Jones, 2002).<br />

The ascospores escape singly through the narrow<br />

pore at the tip of the ascus and usually, but not<br />

invariably, break up in<strong>to</strong> constituent part-spores<br />

outside the ascus.<br />

There is an exceptionally wide range of<br />

conidial forms including the anamorph genera<br />

Paecilomyces, Hirsutella, Hymenostilbe, Beauveria,<br />

Metarhizium and Tolypocladium (Hodge, 2003;<br />

Stensrud et al., 2005). Some of these are<br />

shown in Fig. 12.35. Evidence for the connection<br />

between anamorphs and their Cordyceps teleomorphs<br />

has been obtained in various ways,<br />

e.g. by the development of anamorphs in cultures<br />

derived from ascospores or from hyphal<br />

bodies within a parasitized insect, by the development<br />

of perithecial stromata on insect larvae


CLAVICIPITALES<br />

361<br />

artificially infected with conidia, or by molecular<br />

comparison of gene sequences (Liu et al., 2002).<br />

Teleomorphic states may be rare and some of<br />

the fungi listed above are far better known as<br />

anamorphs than as teleomorphs. This is especially<br />

true of Beauveria bassiana and Metarhizium<br />

anisopliae, both of which are widely distributed<br />

in soil, probably growing as saprotrophs.<br />

They have been used in the biological control of<br />

insect pests.<br />

Cordyceps<br />

Cordyceps militaris forms club-shaped orange- or<br />

red-coloured stromata (Plate 5g) which project<br />

above the ground in autumn from buried lepidopteran<br />

larvae and pupae (Winterstein, 2001).<br />

Species from several genera of Lepidoptera and<br />

some Hymenoptera are susceptible. If ascospore<br />

segments or conidia come in<strong>to</strong> contact with the<br />

integument of a pupa, germination occurs and is<br />

followed by penetration of the cuticle, aided by<br />

the secretion of chitinolytic enzymes. Soon after<br />

penetration, cylindrical hyphal bodies appear<br />

in the haemocoel. The hyphal bodies increase in<br />

number by budding, and become distributed<br />

within the insect’s body. Death of the insect is<br />

probably associated with the secretion of the<br />

<strong>to</strong>xin cordycepin (3’-deoxyadenosine) which also<br />

accumulates in the perithecial stroma (Yu et al.,<br />

2001; Kim et al., 2002). After death some 5 days<br />

after infection, mycelial growth takes place and<br />

the body of the dead insect becomes transformed<br />

in<strong>to</strong> a sclerotium. Under suitable conditions<br />

one or more perithecial stromata develop<br />

above ground, some 45 60 days post infection.<br />

Perithecial stromata can also form in pure<br />

culture on rice grain supplemented with haemoglobin<br />

or casein (Basith & Madelin, 1968). In pure<br />

cultures derived from single ascospores, a phialidic<br />

conidial state called Paecilomyces militaris<br />

is formed (Fig 12.33c).<br />

It is not known whether C. militaris is homoor<br />

heterothallic. Perithecial development has<br />

been studied in C. militaris by Varitchak (1931).<br />

Coiled septate ascogonia arise in the peripheral<br />

layers of the perithecial stroma. The segments<br />

of the ascogonium become multinucleate and<br />

give rise <strong>to</strong> ascogenous hyphae from which asci<br />

develop in a single cluster at the base of the<br />

Fig12.33 Cordyceps militaris. (a) Two perithecial stromata<br />

attached <strong>to</strong> pupae. (b) Ascus and multiseptate ascospores.<br />

Note the ascus cap. (c) Conidiophores and conidia.<br />

perithecium. The perithecial wall is derived<br />

from hyphae which develop from the stalk of<br />

the ascogonium or from surrounding hyphae.<br />

Paraphysis-like hyphae grow inwards from the<br />

perithecial wall, but at maturity these hyphae<br />

dissolve and disappear.<br />

Cordyceps sinensis (Fig. 12.34) is highly prized in<br />

traditional Chinese medicine (Pegler et al., 1994).<br />

It grows on the larvae of hepialid moths at<br />

3600 5000 m in mountainous areas in southern


362 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.34 Stromata of Cordyceps sinensis growing from insect<br />

larvae. Image kindly supplied byY.-J.Yao.<br />

and western China, Tibet and Nepal (Jiang & Yao,<br />

2002). The cylindrical perithecial stromata may<br />

be 10 cm or more in length and bundles of dried<br />

stromata are sold commercially. The anamorph<br />

is Hirsutella sinensis (Liu et al., 2001). The conidiogenous<br />

cells are tightly packed <strong>to</strong>gether on the<br />

surface of a stroma.<br />

The ecological role of Cordyceps spp. on insects<br />

is difficult <strong>to</strong> evaluate. Evans and Samson (1982,<br />

1984), judging by the large numbers of stromata<br />

collected on worker ants in tropical forests, were<br />

of the opinion that they exerted a controlling<br />

effect on population size.<br />

Cordyceps ophioglossoides grows in woods on<br />

the subterranean ascocarps of the hart’s truffle<br />

Elaphomyces, forming bright yellow mycelial<br />

stands over its surface. Brown club-shaped<br />

perithecial stromata grow above ground in<br />

autumn (Plate 4f). Cordyceps capitata grows in<br />

similar situations. Phylogenetic analyses using<br />

nuclear and mi<strong>to</strong>chondrial ribosomal DNA<br />

derived from 22 species of Cordyceps and their<br />

known anamorphs have shown that four species<br />

of Cordyceps parasitic on Elaphomyces have very<br />

close similarity <strong>to</strong> two species which parasitize<br />

the nymphs of cicadas, and have probably<br />

evolved from them, an example of ‘interkingdom<br />

host-jumping’. These related species have been<br />

grouped <strong>to</strong>gether in a ‘truffle cicada clade’ and<br />

it has been suggested that the truffle cicada<br />

clade separated from other Cordyceps clades<br />

about 43 + 13 million years ago (Nikoh &<br />

Fukatsu, 2000). Cicada nymphs develop for<br />

several years underground, feeding on xylem<br />

sap from the host trees with which they are<br />

associated. In this respect their physiology is<br />

similar <strong>to</strong> that of Elaphomyces spp., which are<br />

mycorrhizal partners of trees, obtaining nutrients<br />

from them. Nikoh and Fukatsu (2000) have<br />

speculated that the overlapping niches of the<br />

species of Cordyceps which parasitize cicadas and<br />

those which parasitize hart’s truffles may have<br />

promoted this interkingdom host jumping.<br />

Species of Cordyceps reproduce asexually,<br />

some having more than one anamorphic state,<br />

i.e. with synanamorphs (Hodge, 2003).<br />

Considerable research attention has been<br />

focused on the three anamorphs Beauveria<br />

bassiana, Metarhizium anisopliae and Tolypocladium<br />

inflatum (Figs. 12.35a c). The first two have<br />

potential for the biological control of insect<br />

pests, whereas Tolypocladium inflatum is the<br />

source of the immunosuppressant drug cyclosporin<br />

A (Fig. 12.24a).<br />

Beauveria bassiana<br />

This is the conidial form of Cordyceps bassiana<br />

(Huang et al., 2002) and causes the serious white<br />

muscardine disease of silkworm (Bombyx mori)<br />

larvae, which is a threat <strong>to</strong> silk production. This<br />

species has been known for well over a century<br />

and is of his<strong>to</strong>rical interest because Agostino<br />

Bassi, who studied the fungus around 1835,<br />

proposed the germ theory of disease on the<br />

basis of his results. This preceded by several years<br />

the publications of Robert Koch, who is usually<br />

given credit for the formal proof of pathogenesis<br />

by micro-organisms. Beauveria bassiana is widely<br />

distributed in the soil, usually associated with<br />

diseased insects (Domsch et al., 1980). Species of<br />

Coleoptera, Diptera, Lepidoptera and other<br />

groups, including Arachnida, are covered by<br />

white dusty raised tufts of hyphae bearing


CLAVICIPITALES<br />

363<br />

conidia. The association has had a long his<strong>to</strong>ry. A<br />

worker ant covered with a fungus similar <strong>to</strong> the<br />

present-day B. bassiana has been discovered<br />

embedded in 25 million year-old amber (Poinar<br />

& Thomas, 1984). Beauveria bassiana grows readily<br />

in culture, forming dry conidia and, occasionally,<br />

synnemata. The conidiophores form densely<br />

clustered whorls of conidiogenous cells which<br />

are swollen at the base and extend in<strong>to</strong> a zigzag<br />

shaped rachis forming small, globose, smooth,<br />

hyaline conidia (Fig. 12.35a).<br />

Several myco<strong>to</strong>xins have been obtained<br />

from cultures of B. bassiana including beauvericin,<br />

a cyclic depsipeptide, oosporein and bassianolide,<br />

all of which are <strong>to</strong>xic <strong>to</strong> insect larvae<br />

(see Boucias & Pendland, 1998). Numerous<br />

attempts have been made <strong>to</strong> use commercial<br />

preparations of conidia in the biological control<br />

of insect pests, such as the Colorado beetle<br />

on pota<strong>to</strong>es or the codling moth on apples.<br />

Conidia are either applied alone or, in integrated<br />

control, in conjunction with a chemical insecticide.<br />

However, dependable success in biocontrol<br />

using B. bassiana is still awaited. Whilst it<br />

is possible <strong>to</strong> produce large quantities of inoculum,<br />

this cannot be s<strong>to</strong>red for very long, and<br />

there are also problems in maintaining a<br />

reproducibly high level of biocontrol of insect<br />

pathogens in outdoor situations (Boucias &<br />

Pendland, 1998).<br />

Metarhizium anisopliae<br />

This fungus is the cause of green muscardine<br />

disease of insects. There are three varieties,<br />

var. anisopliae, var. acridum and var. major. The<br />

teleomorph of M. anisopliae var. major is C. brittlebankisoides<br />

(Liu et al., 2001, 2002). Metarhizium<br />

anisopliae grows in soil (Domsch et al., 1980), but<br />

it is also one of the most important insect<br />

pathogens. Metarhizium anisopliae var. anisopliae<br />

has a wide host range, attacking members of the<br />

Coleoptera, Orthoptera, Hemiptera and<br />

Hymenoptera as well as Arachnida. Metarhizium<br />

anisopliae var. major is more host-specific, mostly<br />

infecting soil-inhabiting scarabeid beetles<br />

(Boucias & Pendland, 1998). In culture it grows<br />

slowly, forming columns of green, shortly cylindrical,<br />

uninucleate phialoconidia which are rich<br />

in lipid droplets (Fig. 12.35b). The wall of the<br />

conidium is three-layered and the outermost<br />

layer is highly hydrophobic due <strong>to</strong> impregnation<br />

by a hydrophobin. There have been extensive and<br />

detailed studies of the physiology and enzymology<br />

of germination and penetration by the germ<br />

tubes, mainly carried out by Charnley and St<br />

Leger (1991). The brief account which follows<br />

Fig12.35 Anamorphic states associated with Cordyceps. (a) Beauveria bassiana.(b)Metarhizium anisopliae. (c) Tolypocladium inflatum.<br />

The phialides have a swollen base and a long neck. (a,b) <strong>to</strong> same scale.


364 HYMENOASCOMYCETES: PYRENOMYCETES<br />

draws heavily on the summary in Boucias and<br />

Pendland (1998). Water is required for germination,<br />

which is followed by attachment <strong>to</strong> the host<br />

cuticle by means of an appressorium.<br />

Appressorium development is stimulated by<br />

contact with a hard surface and can take place<br />

on glass or polystyrene, so long as complex<br />

nitrogenous substances (e.g. yeast extract or<br />

pep<strong>to</strong>ne) are present. In nature it is presumed<br />

that such compounds are derived from the<br />

cuticle. The wall of the appressorium is<br />

surrounded by a coat of mucilage which tightly<br />

attaches the appressorium <strong>to</strong> the host integument.<br />

An infection peg from the appressorium<br />

penetrates the surface layer of the integument,<br />

the epicuticle, but when it reaches the procuticle<br />

(the layer beneath it), the infection peg expands<br />

<strong>to</strong> form a penetration plate which grows out<br />

parallel <strong>to</strong> the surface of the integument. From<br />

the penetration plate, penetration hyphal bodies<br />

develop and, from these, vertical penetration<br />

hyphae grow through the innermost layer, the<br />

procuticle, <strong>to</strong> the hypodermis and the body<br />

cavity. From the vertical penetration hyphae,<br />

hyphal bodies in turn develop which become<br />

dispersed in the haemolymph. The hyphal bodies<br />

come <strong>to</strong> rest in the fat bodies of the insect and<br />

then give rise <strong>to</strong> mycelium, by which time,<br />

48 72 h post infection, the host is dead. Death<br />

is probably brought about by the action of<br />

secondary metabolites which function as <strong>to</strong>xins.<br />

These include a series of depsipeptides<br />

(destruxins A E), a hydrophobin, cy<strong>to</strong>chalasins<br />

and alkaloids.<br />

Metarhizium anisopliae var. acridum (formerly<br />

known as M. flavoviride) has been used in the<br />

biological control of grasshoppers and locusts<br />

by suspending conidia in oil for low volume<br />

application. As for B. bassiana, however, the<br />

commercial viability of M. anisopliae as a mainstream<br />

insecticide remains <strong>to</strong> be established<br />

(Hajek et al., 2001).<br />

Tolypocladium inflatum<br />

This is the anamorph of Cordyceps subsessilis which<br />

fruits on the larvae of scarabeid beetles (Hodge<br />

et al., 1996). It is the commercial source of the<br />

immuno-suppressant cyclosporin A, which has<br />

become a crucial drug in the treatment of the<br />

rejection reaction after organ transplantation<br />

(Dreyfuss et al., 1976; Borel, 1986). Other secondary<br />

metabolites are the efrapeptins, compounds<br />

with anti-fungal and insecticidal properties<br />

(Krasnoff & Gupta, 1992). Tolypocladium inflatum<br />

grows in soil. In culture it first forms Acremoniumlike<br />

conidia on single slender phialides, but<br />

later copious hyaline conidia are produced<br />

in slime from clusters of phialides with a<br />

globose base and a long, narrow tapering neck<br />

(Fig. 12.35c).<br />

12.6 Ophios<strong>to</strong>matales<br />

This group contains about 6 genera (110 species)<br />

of perithecial ascomycetes which are mainly<br />

saprotrophic or parasitic on woody hosts. The<br />

perithecia are non-stromatic, generally longnecked<br />

and solitary. We shall consider only<br />

Ophios<strong>to</strong>ma.<br />

12.6.1 Ophios<strong>to</strong>ma<br />

There are about 100 species of Ophios<strong>to</strong>ma (Grylls<br />

& Seifert, 1993; Seifert et al., 1993), largely<br />

confined <strong>to</strong> wood and bark and associated with<br />

bark-boring beetles which disperse their ascospores<br />

and conidia. Some cause fatal diseases<br />

of trees, notably Dutch elm disease (see below).<br />

Others, e.g. O. piceae (Fig. 12.36), cause blue-stain<br />

(¼ sap-stain) of conifer wood (Seifert 1993;<br />

Gibbs, 1993). The anamorphic fungus Sporothrix<br />

schenkii, which is closely related <strong>to</strong> Ophios<strong>to</strong>ma<br />

(Berbee & Taylor, 1992b), is a human pathogen<br />

(Summerbell et al., 1993).<br />

The perithecia of Ophios<strong>to</strong>ma are black in<br />

colour with a bulbous base and a long cylindrical<br />

neck, the ostiole of which is surmounted by<br />

a ring of stiff tapering hairs which hold the<br />

hyaline unicellular ascospores in a mucilaginous<br />

blob. The asci have thin evanescent walls and<br />

the ascospores are released in<strong>to</strong> the body of the<br />

perithecium and move up the narrow tube inside<br />

the neck. The perithecia closely resemble those<br />

of Cera<strong>to</strong>cystis and the two genera are sometimes<br />

synonymized (see de Hoog & Scheffer, 1984).<br />

However, molecular and morphological comparisons<br />

have indicated that, despite their


OPHIOSTOMATALES<br />

365<br />

Fig12.36 Ophios<strong>to</strong>ma piceae.<br />

(a) Perithecium showing spore drop at tip<br />

of neck. (b) Details of ostiole with ring of<br />

setae. (c) Asci. (d) Synnema<strong>to</strong>us<br />

conidiophore. (e) Details of apex of<br />

conidiophore.<br />

similarity, the two genera are not closely related,<br />

Ophios<strong>to</strong>ma being in a sister clade <strong>to</strong> the<br />

Diaporthales (p. 373) whilst Cera<strong>to</strong>cystis is allied<br />

<strong>to</strong> the Microascales (p. 368; Hausner et al., 1993;<br />

Spatafora & Blackwell, 1994). The similarity in<br />

perithecial morphology suggests parallel evolution<br />

in adaptation <strong>to</strong> the insect dispersal of<br />

ascospores (Malloch & Blackwell, 1993). There<br />

are differences in anamorphs between the two<br />

groups. In Cera<strong>to</strong>cystis sensu stric<strong>to</strong> the anamorphs<br />

are phialidic, of the Thielaviopsis type (see<br />

Figs. 8.8, 12.42e), whereas the anamorphs of<br />

Ophios<strong>to</strong>ma are annellidic (de Hoog, 1974). Other<br />

points of difference are listed on p. 369.<br />

Most species of Ophios<strong>to</strong>ma are heterothallic<br />

and the ascospores are of two mating types,<br />

A and B. Although ascogonia have been described,<br />

there are no antheridia. When isolates<br />

of different mating types from the same ascocarp<br />

of O. ulmi or O. piceae are inoculated near each<br />

other in agar cultures, the two mycelia intermingle<br />

and perithecia develop. However, when<br />

cultures of different origin are opposed <strong>to</strong> each<br />

other, a line of barrage analogous <strong>to</strong> that seen<br />

in Podospora anserina may develop between the<br />

approaching mycelia, a phenomenon associated<br />

with vegetative incompatibility (Brasier, 1993).<br />

Perithecium development in O. ulmi has<br />

been described by Rosinski (1961) and the ultrastructure<br />

of the ascogenous hyphae and ascosporogenesis<br />

by Jeng and Hubbes (1980). The<br />

multinucleate coiled ascogonium, differentiated<br />

by its greater width from the hyphae which<br />

subtend it, becomes surrounded by a mantle<br />

of branched hyphae probably derived from the<br />

ascogonial stalk. Ascogenous hyphae arise as<br />

buds from the ascogonium which is positioned<br />

near the base of the ascocarp. A central cavity<br />

arises as the result of enlargement of the outermost<br />

cells of the ascocarp, and the ascogenous


366 HYMENOASCOMYCETES: PYRENOMYCETES<br />

hyphae form a lining layer around this cavity<br />

and develop centripetally <strong>to</strong>wards the centre of<br />

the cavity. Croziers develop at the tips of the<br />

ascogenous hyphae and produce a succession<br />

of asci, indicating the ascohymenial nature of<br />

the fungus. The asci have extremely thin walls.<br />

There are no paraphyses and no periphyses.<br />

Ascospores released in<strong>to</strong> the cavity of the<br />

perithecium are extruded through the narrow<br />

neck and accumulate in a mucilaginous blob<br />

held in place by the ring of ostiolar hairs. Barkboring<br />

beetles which feed on the ascospores<br />

(and also conidia) help in their dispersal. The<br />

beetles are attracted by ‘fruity’ odours emitted<br />

from the mycelium. These are volatile metabolites,<br />

mainly short-chain alcohols and esters, as<br />

well as monoterpenes and sesquiterpenes<br />

(Hanssen, 1993).<br />

Some species of Ophios<strong>to</strong>ma have two or<br />

more synanamorphs. These vary in structure<br />

from yeast-like <strong>to</strong> mononema<strong>to</strong>us or synnema<strong>to</strong>us.<br />

They have been assigned <strong>to</strong> several<br />

anamorph genera including Lep<strong>to</strong>graphium,<br />

Sporothrix, Pesotum and Graphium. For example,<br />

O. ulmi has a yeast-like anamorph, a mononema<strong>to</strong>us<br />

anamorph referred <strong>to</strong> Pesotum and a synnema<strong>to</strong>us<br />

anamorph, Graphium ulmi. Older hyphae<br />

may also produce endoconidia (Fig. 12.37).<br />

Wingfield et al. (1991) have placed Pesotum in<br />

synonymy with Graphium.<br />

12.6.2 Dutch elm disease<br />

Dutch elm disease is a vascular wilt disease<br />

of Ulmus spp. caused by O. ulmi, O. novo-ulmi and<br />

O. himal-ulmi. It is best regarded as a disease<br />

complex because it is invariably associated with<br />

the activities of bark-boring scolytid beetles such<br />

as Scolytus scolytus, S. multistriatus and Hylurgopinus<br />

rufipes. These are the vec<strong>to</strong>rs for the disease.<br />

Elm populations worldwide have been ravaged,<br />

causing the death of millions of trees in Europe<br />

and North America and changing the appearance<br />

of the landscape, particularly where hedgerow<br />

elms have been killed. Diseased tree leaves<br />

wilt in dry weather and rapidly turn brown<br />

and brittle. Defoliation and death of the twigs<br />

ensues, and eventually the whole tree dies<br />

(Plate 5h). The fungus persists as a saprotroph<br />

in the bark of dead trees. Infected twigs show<br />

a characteristic brown flecking in the sapwood.<br />

This is associated with the development of<br />

brown-coloured bladder-like inflated cells of the<br />

xylem parenchyma called tyloses, which invade<br />

the xylem vessels and block them. Gums released<br />

partly by the action of cell wall-degrading enzymes<br />

of the pathogen impede water flow. Wilting<br />

is also associated with the production of the<br />

wilt <strong>to</strong>xin cera<strong>to</strong>-ulmin, which is a hydrophobin<br />

(Richards, 1993). However, since mutants of<br />

O. novo-ulmi with low cera<strong>to</strong>-ulmin production<br />

still retain pathogenicity, an alternative role<br />

for it has been sought. Temple et al. (1997) have<br />

shown that cera<strong>to</strong>-ulmin can enhance the adhesiveness<br />

of yeast-like propagules of the pathogen<br />

and also protect them from desiccation.<br />

Dutch elm disease was first described in<br />

Holland in the late 1920s and spread <strong>to</strong> the rest<br />

of north-western Europe, North America and <strong>to</strong><br />

parts of Asia. The disease declined in severity<br />

in Europe in the 1940s but persisted in North<br />

America, possibly because the American elms<br />

were more susceptible. This first pandemic was<br />

relatively mild in effect and did not destroy<br />

the elm tree populations. When first discovered,<br />

the disease was associated with the synnematal<br />

conidial state, Graphium ulmi. Later the<br />

teleomorphic state was discovered and named<br />

Cera<strong>to</strong>s<strong>to</strong>mella ulmi, then Cera<strong>to</strong>cystis ulmi, now<br />

O. ulmi. In the mid-1960s simultaneous outbreaks<br />

of a more severe and aggressive form of the<br />

disease occurred, centred around ports in<br />

southern England and originating from<br />

infected elm logs from North America. These<br />

outbreaks spread rapidly from the original<br />

infection foci in<strong>to</strong> much of Britain and mainland<br />

Europe.<br />

Isolations from trees affected by the aggressive<br />

form of the disease yielded a strain of<br />

Ophios<strong>to</strong>ma distinguished by its fluffy appearance<br />

and more rapid growth in culture, in contrast<br />

with the waxy appearance and slower growth<br />

of the non-aggressive strain. In many places<br />

where O. ulmi was present, it has now been<br />

replaced by the more aggressive form (Brasier<br />

et al., 1998). The aggressive strain is regarded as<br />

a distinct species, O. novo-ulmi (Brasier, 1991b).<br />

Closer investigation of isolates of O. novo-ulmi,


OPHIOSTOMATALES<br />

367<br />

Fig12.37 Asexual reproduction in Ophios<strong>to</strong>ma ulmi. (a,b) Synnema<strong>to</strong>us conidiophores (Graphium ulmi). Parallel bundles of dark<br />

hyphae branch at their tips <strong>to</strong> produce holoblastic conidia which accumulate in a sticky drop. (c,d) Mononema<strong>to</strong>us synanamorph<br />

terminating in conidiogenous cells with a succession of holoblastic conidia which, on detachment, leave a protruding scar or denticle.<br />

In d some of the conidia show yeast-like budding. (e) An older hypha within which a new hypha and endoconidia have developed<br />

(from Harris & Taber,1973).<br />

based on their vegetative incompatibility characteristics,<br />

showed that they can be separated<br />

in<strong>to</strong> two distinct vc (vegetative compatibility)<br />

supergroups. One biotype, centred on the<br />

Romania Moldova Ukraine region of Europe,<br />

was designated the Eurasian or EAN vc supergroup,<br />

whilst the other, centred on the Southern<br />

Great Lakes region of North America, has<br />

been designated the North American or NAN vc<br />

supergroup. They have been formally recognized<br />

as distinct subspecies, O. novo-ulmi subsp. novoulmi<br />

for the EAN strains and subsp. americana<br />

for the NAN strains. There are morphological<br />

differences between the perithecia of the two<br />

subspecies, those of subspecies novo-ulmi having<br />

longer necks than those of subspecies americana<br />

(Brasier & Kirk, 2001). Naturally occurring<br />

hybrids between the two subspecies have been<br />

detected. Rare interspecific hybrids between<br />

O. ulmi and O. novo-ulmi also occur where the<br />

former species is being replaced by the latter<br />

(Brasier et al., 1998). Hybridization between introduced<br />

pathogens permits their rapid evolution<br />

(Brasier, 2001). Ophios<strong>to</strong>ma himal-ulmi is endemic<br />

<strong>to</strong> the Himalayas and has been distinguished as<br />

a third species (Brasier & Mehrotra, 1995).<br />

Fertilized females of bark-boring beetles,<br />

possibly carrying ascospores or conidia, excavate<br />

tunnels beneath the bark of living trees, often<br />

those already weakened by the disease, <strong>to</strong> lay<br />

a cluster of eggs. After hatching, the developing<br />

larvae also make tunnels beneath the bark<br />

radiating outwards from the egg chamber. They<br />

feed on the infected wood, and the galleries<br />

which they excavate are often lined by synnemata<br />

or perithecia, so that conidia and ascospores<br />

become attached <strong>to</strong> their mouthparts and<br />

bodies. Young, sexually immature adults emerge<br />

in the spring and summer. They fly <strong>to</strong> the<br />

crotches (branch points) of young twigs where<br />

they feed on bark before maturation and<br />

mating. During this twig-feeding stage spores<br />

of the fungal pathogen may be introduced in<strong>to</strong><br />

the host sapwood, and from these multiple<br />

inoculation points the fungus may spread in<strong>to</strong><br />

host tissues by mycelial growth or by movement


368 HYMENOASCOMYCETES: PYRENOMYCETES<br />

of conidia in xylem vessels (Webber & Brasier,<br />

1984; Webber & Gibbs, 1989). By the latter<br />

method it has been estimated that movement<br />

can be as much as 10 cm day 1 . In addition <strong>to</strong><br />

being transmitted by insect vec<strong>to</strong>rs, the pathogen<br />

can be passed from tree <strong>to</strong> tree by natural<br />

root contact.<br />

The control of Dutch elm disease has been<br />

attempted by the combined use of fungicides and<br />

insecticides but is now based largely on the<br />

breeding of resistant cultivars. Major genes for<br />

resistance have been identified in a group of<br />

Asian species. By crossing some of these with<br />

European elms, cultivars have been bred and<br />

released for commercial sale (Smalley & Guries,<br />

1993). Resistance may be correlated with the<br />

production of phy<strong>to</strong>alexins (mansonones) by the<br />

host in response <strong>to</strong> infection (Smalley et al., 1993).<br />

In U. minor a correlation has also been found<br />

between vessel diameter and susceptibility, trees<br />

having vessels of large diameter being more<br />

susceptible <strong>to</strong> the disease than those with<br />

smaller diameter vessels (Solla & Gil, 2002). An<br />

interesting novel potential method of control<br />

is by the use of hypovirulent strains of the<br />

pathogen infected by cy<strong>to</strong>plasmically transmissible<br />

virus-like agents called d-fac<strong>to</strong>rs (d for<br />

disease), now known <strong>to</strong> consist of doublestranded<br />

mi<strong>to</strong>chondrial RNA elements. Twelve<br />

d-fac<strong>to</strong>rs have been characterized. Strains<br />

carrying d-fac<strong>to</strong>rs are hypovirulent and this<br />

is correlated with a reduced capacity in vitro<br />

<strong>to</strong> produce cera<strong>to</strong>-ulmin (Rogers et al., 1986;<br />

Sutherland & Brasier, 1995).<br />

12.7 Microascales<br />

The Microascales are a small order currently<br />

containing 67 species (Kirk et al., 2001).<br />

The taxonomy of the Microascales has seen a<br />

turbulent his<strong>to</strong>ry. Species accommodated here<br />

are characterized by perithecia (rarely cleis<strong>to</strong>thecia)<br />

which usually have a long neck. The fact that<br />

asci are scattered throughout the perithecial<br />

cavity and are not produced by croziers is one<br />

reason why members of the Microascales have<br />

been considered in the past <strong>to</strong> belong <strong>to</strong> the<br />

Plec<strong>to</strong>mycetes. Another moot point has been the<br />

great similarity of perithecia and general ecological<br />

features between Cera<strong>to</strong>cystis (Microascales)<br />

and Ophios<strong>to</strong>ma (Ophios<strong>to</strong>matales), which will be<br />

discussed in more detail below. In its current<br />

shape, the order Microascales is monophyletic by<br />

DNA analyses (Hausner et al., 1993). Because the<br />

two major lineages, the Cera<strong>to</strong>cystidaceae and<br />

Microascaceae, show considerable differences in<br />

their biology and ecology, we will briefly discuss<br />

members of both groups.<br />

12.7.1 Microascus (Microascaceae)<br />

The family Microascaceae currently contains<br />

43 species in 8 genera. The conidial forms are<br />

hyphomyce<strong>to</strong>us, with conidia produced on<br />

annellides (see Fig. 8.7). The most important<br />

genus is Microascus with 14 species, which have<br />

been described in detail by Barron et al. (1961).<br />

Simple conidiogenous forms of Microascus are<br />

referable <strong>to</strong> Scopulariopsis (Fig. 12.38), whereas<br />

more complex, synnema<strong>to</strong>us forms belong<br />

<strong>to</strong> Cephalotrichum (formerly called Dora<strong>to</strong>myces;<br />

Fig. 12.39) or Trichurus (Fig. 12.40). In agar culture,<br />

both Cephalotrichum and Trichurus produce<br />

simple Scopulariopsis-like forms in addition <strong>to</strong><br />

the striking synnemata. The connection between<br />

these annellidic forms and Microascus has<br />

now been confirmed by phylogenetic analyses<br />

(Issakainen et al., 2003). Just <strong>to</strong> confuse matters,<br />

Cephalotrichum stemonitis produces, in addition <strong>to</strong><br />

annelloconidia, a distinct Echinobotryum conidial<br />

state (Fig. 12.39e). Graphium is another formgenus<br />

producing synnemata which form annelloconidia<br />

at their apex (Fig. 12.41), but here the<br />

conidia are held in a drop of mucilage at the<br />

tip of the synnema. Certain Graphium-like forms,<br />

e.g. G. penicillioides (Fig. 12.41a), have affinity with<br />

Microascus, although others seem <strong>to</strong> belong <strong>to</strong><br />

the Ophios<strong>to</strong>matales (Okada et al., 2000).<br />

Microascus spp. and members of related<br />

genera (e.g. Pseudoallescheria, Petriella) are capable<br />

of metabolizing a wide range of carbon sources,<br />

including keratin, crude oil, cellulose and even<br />

phenol, and they are <strong>to</strong>lerant of wide-ranging<br />

environmental conditions. Consequently, they<br />

are found from the Arctic <strong>to</strong> hot desert soil, in<br />

salt marshes, bat guano, herbivore dung, food,


MICROASCALES<br />

369<br />

Fig12.38 Scopulariopsis brevicaulis. Conidiophores terminating<br />

in conidiogenous cells (annellides) from which chains of conidia<br />

develop.The stippled areas at the tips of the annellides indicate<br />

the region of growth associated with the development of<br />

successive conidia.<br />

silage, and on keratin-rich substrates. Because<br />

of their ability <strong>to</strong> degrade keratin and grow at<br />

37°C, some Scopulariopsis spp. can be pathogenic<br />

<strong>to</strong> humans, occasionally causing deep-seated<br />

mycoses (de Hoog et al., 2000a). Not surprisingly<br />

for such versatile and adaptable fungi, they<br />

quickly develop resistance against commonly<br />

used antifungal drugs (Cuenca-Estrella et al.,<br />

2003). This feature, rather than their pathogenicity<br />

per se, seems <strong>to</strong> be the main reason why<br />

they can be troublesome in clinical situations.<br />

Several well-known Scopulariopsis states<br />

have only recently been assigned <strong>to</strong><br />

heterothallic Microascus species (Abbott & Sigler,<br />

2001). Detailed developmental studies of perithecium<br />

formation in homothallic species have<br />

been carried out by Corlett (1966). The vegetative<br />

hyphae of Microascus consist of uninucleate segments.<br />

Sexual reproduction is initiated when<br />

a coiled ascogonium forms and becomes<br />

ensheathed by hyphae arising from the ascogonial<br />

base as well as the surrounding vegetative<br />

mycelium. At this primordium stage, the outermost<br />

layer already becomes melanized, and it<br />

surrounds the developing pseudoparenchyma.<br />

The primordium enlarges by further growth of<br />

the pseudoparenchyma. The ascogonium is positioned<br />

initially in a cavity in the centre of the<br />

primordium, but soon sterile hyphae grow<br />

inwards from the surrounding pseudoparenchyma,<br />

raising the ascogonium <strong>to</strong>wards the<br />

apex of the perithecium. From the ascogonium,<br />

ascogenous hyphae grow between the sterile,<br />

inwardly radiating hyphae and form asci without<br />

croziers from their tips which have become<br />

binucleate. Simultaneously, at the tip of the<br />

perithecium pseudoparenchyma<strong>to</strong>us hyphae<br />

begin <strong>to</strong> grow inwards and then upwards,<br />

rupturing the apical wall and extending a<br />

perithecial neck with an ostiole. During maturation<br />

of the perithecium, the sterile hyphae in<br />

the perithecial centre lyse, followed by disintegration<br />

of the ascus walls, so that the mature<br />

perithecium contains an outer melanized<br />

layer surrounding a pseudoparenchyma which<br />

encloses masses of free ascospores. These ooze<br />

out through the ostiole as a tendril.<br />

12.7.2 Cera<strong>to</strong>cystis and Sphaeronaemella<br />

(Cera<strong>to</strong>cystidaceae)<br />

The most important genus of this small family<br />

is Cera<strong>to</strong>cystis (14 species) associated with living<br />

trees. Some species cause vascular wilts, perhaps<br />

the most important pathogen being C. fagacearum,<br />

the cause of oak wilt, which is especially<br />

destructive in the United States. Cera<strong>to</strong>cystis<br />

fimbriata attacks an exceptionally wide range of<br />

plants, including the European plane (Platanus<br />

acerifolia), Eucalyptus spp., mango, coffee, sweet<br />

pota<strong>to</strong> and rubber (Kile, 1993). However, infections<br />

by Cera<strong>to</strong>cystis spp. need not be destructive<br />

<strong>to</strong> the host, and some species produce benign


370 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.39 Cephalotrichum (Dora<strong>to</strong>myces) stemonitis. (a) Synnema or macronema<strong>to</strong>us conidiophore consisting of a parallel bundle<br />

of dark hyphae branching at their tips <strong>to</strong> form conidiogenous cells (annellides) and chains of conidia. (b) Developing synnema.<br />

(c) Micronema<strong>to</strong>us conidiophore bearing six annellides.The stippled region is the cylinder of annellide growth. (d) Annellides from<br />

a synnema showing chains of annelloconidia.The stippled zone represents the growth cylinder of the annellide. (e) Echinobotryum<br />

conidial state. (b) and (e) <strong>to</strong> same scale, (c) and (d) <strong>to</strong> same scale.


MICROASCALES<br />

371<br />

Fig12.40 Trichurus sp. on agar. (a) Macronema<strong>to</strong>us<br />

fructification. Note the melanized curved setae typical of the<br />

form-genus.Conidia are produced from annellides arising<br />

from the tips of the parallel hyphae making up the synnema.<br />

(b) Micronema<strong>to</strong>us fructification with annellidic<br />

conidiogenesis.<br />

infections which merely stain the wood. This<br />

so-called sap-stain can none the less be of<br />

economic importance as it reduces the market<br />

value of the infected wood.<br />

The asci of the Cera<strong>to</strong>cystidaceae (Fig. 12.42c)<br />

deliquesce early in the development of the<br />

perithecium. The perithecium has a very long<br />

neck (Fig. 12.42a) through which the ascospores<br />

are exuded single file, accumulating in a sticky<br />

drop at the tip of the perithecial neck<br />

(Fig. 12.42b). This is thought <strong>to</strong> be an adaptation<br />

<strong>to</strong> dispersal by insects which may pick up the<br />

slime containing ascospores upon browsing,<br />

or by feeding directly on it. Many Cera<strong>to</strong>cystis<br />

spp. produce fruity smells (pineapple, banana,<br />

pear) which attract insects. The substances responsible<br />

are complex mixtures of alcohols, esters<br />

and other volatile substances (Soares et al., 2000),<br />

and industrial processes are being developed <strong>to</strong><br />

produce them under fermentation conditions<br />

(Bluemke & Schrader, 2001). The conidial forms<br />

of Cera<strong>to</strong>cystis are phialidic, and the conidia<br />

are typically barrel-shaped or cylindrical and<br />

are produced deep inside the phialide (Figs. 8.8,<br />

12.42e). This anamorph used <strong>to</strong> be called Chalara<br />

but has now been renamed Thielaviopsis (Paulin-<br />

Mahady et al., 2002). Additionally, dark melanized<br />

chlamydospores may be present.<br />

In addition <strong>to</strong> Cera<strong>to</strong>cystis spp., which are<br />

mainly associated with trees, other members of<br />

this group are soil-borne saprotrophs or weak<br />

pathogens attacking the roots of herbaceous<br />

plants. Examples are Thielaviopsis basicola and<br />

T. thielavioides (Fig. 12.42e), which can cause<br />

significant post-harvest rots in s<strong>to</strong>red carrots<br />

(Punja et al., 1992). The genus Sphaeronaemella<br />

probably also belongs <strong>to</strong> the Cera<strong>to</strong>cystidaceae.<br />

Its perithecium is typical of the family in being<br />

long-necked and <strong>to</strong>pped by a frill of hyphae<br />

which hold the ascospore drop in place<br />

(Figs. 12.42a,b). Sphaeronaemella fimicola grows<br />

on dung, and its distinctive perithecia are<br />

commonly found in association with the fructifications<br />

of other coprophilous fungi. Although<br />

S. fimicola can be a weak hyphal parasite, it<br />

also seems <strong>to</strong> require diffusible metabolites and,<br />

in turn, supplies other metabolites <strong>to</strong> fellow<br />

coprophilous fungi (Weber & Webster, 1998b).<br />

There is also an interaction of S. fimicola with<br />

animals because its ascospore drop appears<br />

<strong>to</strong> stick <strong>to</strong> mites brushing past the perithecial<br />

neck, and the mites in turn hijack rides on<br />

flies (Malloch & Blackwell, 1993). In contrast,<br />

S. helvellae is a mycoparasite infecting the fruit<br />

bodies of Helvella spp. The conidial state of<br />

Sphaeronaemella is called Gabarnaudia and<br />

differs from Thielaviopsis mainly in that the<br />

conidia are formed at the tip of the phialide<br />

(Fig. 12.42d), not inside.


372 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.41 Scanning EM of synnemata of Graphium.The annellations are prominently visible. (a) Graphium penicillioides.<br />

(b) Graphium sp. Both images reprinted from Okada et al. (2000) with kind permission of Centraalbureau voor Schimmelcultures<br />

(Utrecht).Original prints kindly provided by G.Okada.<br />

Fig12.42 Cera<strong>to</strong>cystidaceae. (a) Perithecium of Sphaeronaemella fimicola. Note the extremely long neck and the crown of ostiolar<br />

hairs. (b) Enlargement of a mature perithecium apex, with the ostiolar hairs supporting a mucilaginous blob containing ascospores.<br />

The mucilage has a frothy appearance. Arrows indicate two ascospores ascending in the neck canal. (c) Ascus and conidium of<br />

S. fimicola. (d) The Gabarnaudia state of S. fimicola.The conidia are delimited at the apex of the phialide. (e) Phialide of Thielaviopsis<br />

basicola. Note that the cylindrical conidia are delimited deep inside the phialide neck. (d,e) <strong>to</strong> same scale.


DIAPORTHALES<br />

373<br />

Aspects of pathogenicity<br />

Although Cera<strong>to</strong>cystis spp. are spread by insects,<br />

they are not normally associated with the<br />

tunnels of bark-boring beetles. Instead, they<br />

infect through pruning wounds or the bark,<br />

and may be introduced by sap-feeding insects.<br />

Transmission from infected <strong>to</strong> adjacent healthy<br />

trees may also occur via root contact. Cera<strong>to</strong>cystis<br />

fagacearum causes oak wilt symp<strong>to</strong>ms by invading<br />

the xylem and blocking the conducting<br />

vessels with mycelium and a gel- or gum-like<br />

substance. Additionally, <strong>to</strong>xins are produced by<br />

C. fagacearum and other wilt-causing species,<br />

and these can reproduce many wilt symp<strong>to</strong>ms<br />

in the absence of the fungus (Pazzagli et al., 1999).<br />

Upon death of the host, C. fagacearum forms<br />

compact mycelial mats underneath the bark.<br />

The hyphae of the mycelial mat produce conidia<br />

which are dispersed by beetles attracted by the<br />

fruity smell. The fungus is heterothallic and the<br />

conidia double up as spermatia. When spermatia<br />

of compatible mating types are carried <strong>to</strong> a<br />

mycelial mat, perithecia are formed.<br />

There are several Cera<strong>to</strong>cystis spp. which are<br />

associated with coniferous trees, partly as pathogens<br />

but also as sap-staining fungi (Harring<strong>to</strong>n &<br />

Wingfield, 1998). These belong <strong>to</strong> a phylogenetically<br />

clearly defined group around C. coerulescens<br />

(Witthuhn et al., 1998). Because conifers are<br />

unusual hosts for Cera<strong>to</strong>cystis and all known<br />

species associated with them are closely related,<br />

it is assumed that Cera<strong>to</strong>cystis has only recently<br />

switched hosts from broad-leaved trees.<br />

Cera<strong>to</strong>cystis versus Ophios<strong>to</strong>ma<br />

Much confusion has arisen in the past between<br />

the genera Cera<strong>to</strong>cystis and Ophios<strong>to</strong>ma which<br />

are now known not <strong>to</strong> be closely related phylogenetically<br />

(Spatafora & Blackwell, 1994).<br />

Alexopoulos et al. (1996) have summarized the<br />

main differences, which are as follows. (1) The<br />

anamorphic state of Ophios<strong>to</strong>ma is annellidic<br />

(Graphium; see Fig. 12.41) whereas that of<br />

Cera<strong>to</strong>cystis and its allies is phialidic<br />

(Thielaviopsis, Gabarnaudia; Fig. 12.42). (2)<br />

Ophios<strong>to</strong>ma spp. infect through insect tunnels<br />

whereas Cera<strong>to</strong>cystis spp. infect through wounds.<br />

(3) The cell wall of Ophios<strong>to</strong>ma spp. contains<br />

rhamnose-based polymers and cellulose, but<br />

neither kind is found in Cera<strong>to</strong>cystis. (4)<br />

Ophios<strong>to</strong>ma spp. are insensitive <strong>to</strong> cycloheximide<br />

whereas Cera<strong>to</strong>cystis spp. are sensitive.<br />

12.8 Diaporthales<br />

The order Diaporthales currently includes some<br />

447 species in 94 genera (Kirk et al., 2001) and<br />

is well separated phylogenetically from the<br />

other orders of perithecial ascomycetes (Zhang<br />

& Blackwell, 2001). Several clades can be resolved<br />

within the Diaporthales, although the naming of<br />

families is problematic at present (Castlebury<br />

et al., 2002). Members of the Diaporthales grow<br />

mainly in the bark of trees as saprotrophs or<br />

parasites. Discula destructiva infects all aboveground<br />

organs of dogwood (Cornus spp.) causing<br />

anthracnose (leaf blight, twig dieback, stem<br />

cankers) which is devastating the native Cornus<br />

populations of North America (Redlin, 1991).<br />

Two other important plant-pathogenic genera<br />

are Diaporthe and Cryphonectria, and these will be<br />

considered in more detail below.<br />

Members of the order Diaporthales are<br />

characterized by perithecia which are produced<br />

in clusters or stromata, often embedded in the<br />

host tissue with their long necks protruding<br />

beyond the surface. The asci are unitunicate<br />

and contain eight ascospores which have one or<br />

more septa. Diaporthales may be homothallic or<br />

heterothallic, in the latter case with a bipolar<br />

mating type system. The conidia are produced by<br />

phialide-like structures lining the inside surface<br />

of pycnidia, i.e. anamorphic Diaporthales were<br />

formerly classified as coelomycetes. Pycnidia<br />

are usually dark-walled (melanized) and have<br />

one or more openings through which the<br />

conidia exude in slimy drops. They are more<br />

commonly encountered than the sexual state<br />

(see Fig. 12.43).<br />

12.8.1 Diaporthe and its anamorph<br />

Phomopsis<br />

Developmental aspects<br />

Perhaps the only detailed developmental study<br />

in Diaporthe was carried out by Jensen (1983) on<br />

D. phaseolorum var. sojae, which is homothallic


374 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.43 Phomopsis phaseoli.(a)Surfaceviewofa<br />

conidium-producing stroma formed in agar culture.<br />

(b) Phialides producing b-conidia. (c) Ovoid a-conidium and<br />

elongated b-conidium. Both types of conidium are produced<br />

from similar phialides. (b) and (c) <strong>to</strong> same scale.<br />

like all species of Diaporthe examined <strong>to</strong> date.<br />

In culture, the mycelium consists of narrow<br />

hyphae containing 3 4 nuclei per segment, and<br />

wider ones with up <strong>to</strong> 15 nuclei per segment.<br />

Hyphae aggregate and swell <strong>to</strong> form a pseudoparenchyma<strong>to</strong>us<br />

stroma which produces both<br />

pycnidia and perithecia. In nature, the pycnidial<br />

Phomopsis state is produced earlier than the<br />

Diaporthe state which is most often seen on<br />

dead plant material. Conidial locules form in the<br />

upper region of the stroma (ec<strong>to</strong>stroma) when<br />

two opposing palisades of hyphae press against<br />

each other, accompanied by lysis of hyphae<br />

bordering the developing slit. In consequence,<br />

the slit becomes convoluted and lined<br />

by hymenium. Ec<strong>to</strong>stromatic cells proliferate <strong>to</strong><br />

form a neck, resulting in the typical shape of the<br />

pycnidium consisting of a more or less globose<br />

structure within which the lobed or folded<br />

hymenium is located, and one or several necks<br />

through which conidia are exuded (Fig. 12.43a).<br />

The conidiogenous cells are interpreted as<br />

phialides. They are awl-shaped, 20 mm long and<br />

tapering from 2 3 mm at their base <strong>to</strong> 1 mm at<br />

their apex (Fig. 12.43b). Two types of conidia are<br />

produced, often within the same conidioma<br />

(Fig. 12.43c). The ovoid a-conidia contain two<br />

lipid droplets and germinate readily in culture,<br />

whereas b-conidia are highly elongated and do<br />

not usually germinate in P. phaseoli or other<br />

Phomopsis spp. They are therefore interpreted as<br />

spermatia (Jensen, 1983).<br />

According <strong>to</strong> Uecker (1988), Phomopsis typically<br />

has a dark stroma producing a- and b-conidia,<br />

the teleomorph being Diaporthe. Unfortunately,<br />

few typical Phomopsis species exist since many<br />

produce only either a- or b-conidia, and/or<br />

lack the Diaporthe state. Further, there are few<br />

distinguishing features in conidial shape<br />

between different Phomopsis species, and species<br />

identification is based mainly on the host species<br />

with which a given strain is associated. This<br />

poses problems because the delimitations of host<br />

ranges are not precisely characterized. Thus, the<br />

taxonomy of Phomopsis is in a state of confusion,<br />

with many synonyms probably in existence.<br />

Uecker (1988) has compiled over 800 Phomopsis<br />

names in current use, but Kirk et al. (2001)<br />

have estimated that only about 100 species of<br />

Phomopsis exist. The genus is thus in urgent need<br />

of an up-<strong>to</strong>-date monographic treatment, Grove<br />

(1935) having been the last mycologist <strong>to</strong> rise<br />

<strong>to</strong> this formidable task.<br />

In D. phaseolorum, sexual reproduction is<br />

initiated by the formation of ascogonial coils in<br />

the lower region of the stroma, termed en<strong>to</strong>stroma<br />

by Jensen (1983). These coils become


DIAPORTHALES<br />

375<br />

enveloped by hyphae which proliferate <strong>to</strong> form<br />

the wall (peridium) of the perithecium, and<br />

by others which form the pseudoparenchyma<strong>to</strong>us<br />

centrum and paraphyses. A neck lined by<br />

periphyses is formed relatively early in perithecium<br />

development. The ascogenous hyphae<br />

arising from the ascogonial coils form a bowlshaped<br />

hymenium in the base of the perithecium.<br />

Shortly before maturity of the<br />

perithecium, the ascogenous hyphae produce<br />

croziers. Karyogamy, meiosis and mi<strong>to</strong>sis all<br />

occur in the usual way, giving rise <strong>to</strong> eight<br />

nuclei which divide once more, so that eight<br />

bicellular ascospores are produced in each ascus.<br />

A typical feature of Diaporthe is that the asci often<br />

become detached from the hymenium before or<br />

after the ascospores are ripe, so that the cavity of<br />

the perithecium is filled with loose asci or free<br />

ascospores. Ascospores are usually discharged<br />

non-violently as sticky tendrils exuding from the<br />

ostiole of the perithecium. The ascus has a<br />

prominent apical apparatus which stains with<br />

iodine.<br />

Ecology<br />

Species of Phomopsis and Diaporthe cause serious<br />

plant diseases of commercial significance,<br />

such as pod and stem blight of soybeans<br />

and other pulse fruits (D. phaseolorum, anamorph<br />

P. phaseoli) or stem canker and leaf necrosis of<br />

sunflower (D. helianthi, anamorph P. helianthi).<br />

Many Phomopsis diseases are caused by species<br />

complexes, i.e. different Phomopsis spp. can be<br />

isolated from the same diseased plant in the<br />

field. All species associated with a given disease<br />

may not be equally pathogenic. A good example<br />

is dead arm of vines, which is caused primarily<br />

by P. viticola. Several other Phomopsis species<br />

which can also be isolated from vines are only<br />

weakly pathogenic or entirely non-pathogenic<br />

(Mostert et al., 2001). <strong>Fungi</strong> colonizing the living<br />

plant host as permanently asymp<strong>to</strong>matic infections<br />

are termed endophytes. The example of<br />

P. viticola illustrates the diffuse boundary<br />

between endophytism and parasitism; some<br />

Phomopsis strains are entirely endophytic whereas<br />

others initially cause latent infections but later<br />

become pathogenic (Mostert et al., 2000). Another<br />

example is provided by Diaporthe <strong>to</strong>xica<br />

(anamorph formerly called P. lep<strong>to</strong>stromiformis)<br />

which infects living lupins as coralloid hyphae<br />

with a limited spread beneath the cuticle.<br />

Such latent infections can persist for many<br />

months until death of the colonized host<br />

organs occurs by natural causes, e.g. senescence<br />

at the end of the growing season. Within<br />

1 2 days of host death, large-scale colonization<br />

of the infected tissue occurs (Shankar et al., 1998).<br />

Colonization by saprotrophic and pathogenic<br />

Phomopsis spp. can be accompanied by the secretion<br />

of large amounts of cell wall-degrading<br />

enzymes which indiscriminately macerate the<br />

plant tissue (Heller & Gierth, 2001). This accounts<br />

for the strongly necrotic nature of many<br />

Phomopsis diseases. Toxins may also be produced<br />

by phy<strong>to</strong>pathogenic species, and these may facilitate<br />

rapid colonization by killing host cells and<br />

preventing an immune response. The Phomopsis<br />

state of D. <strong>to</strong>xica produces phomopsins, which are<br />

cyclic peptides comprising six unusual amino<br />

acids. These can accumulate in lupin stems and<br />

seeds <strong>to</strong> such high levels that they cause a serious<br />

<strong>to</strong>xicosis called lupinosis in sheep grazing on<br />

lupin stubble, or fed with lupin seeds (Culvenor<br />

et al., 1977). They seem <strong>to</strong> act mainly on the<br />

microtubular cy<strong>to</strong>skele<strong>to</strong>n, and their primary<br />

effect is on the liver of affected sheep (Edgar<br />

et al., 1986). Another example is the <strong>to</strong>xin phomozin<br />

which is produced by P. helianthi and has<br />

been shown <strong>to</strong> be capable of killing host tissue<br />

(Mazars et al., 1991).<br />

Recently, Diaporthe ambigua, the cause of<br />

cankers on roots and stems of fruit trees,<br />

has been found <strong>to</strong> be infected by a mycovirus<br />

which reduces the ability of its fungal host <strong>to</strong><br />

cause disease on plants (Preisig et al., 2000).<br />

Hypovirulence-causing fungal viruses and their<br />

implications for biological control are discussed<br />

in the following section on Cryphonectria<br />

parasitica.<br />

12.8.2 Cryphonectria parasitica<br />

Readable accounts of various aspects of<br />

C. parasitica have been written by Nuss (1992),<br />

Heiniger and Rigling (1994) and Dawe and<br />

Nuss (2001). The origin of C. parasitica is uncertain.<br />

It was first reported in 1904 as a sudden and


376 HYMENOASCOMYCETES: PYRENOMYCETES<br />

rapidly spreading infection of Castanea dentata<br />

(American edible chestnut) in North America,<br />

and in 1938 on C. sativa (European edible chestnut)<br />

in Italy, from where it spread rapidly <strong>to</strong><br />

other countries. It is also known in Asia, the<br />

putative centre of origin of C. sativa which was<br />

brought from the Near East <strong>to</strong> Western Europe<br />

by the Romans. It is possible that the pathogen<br />

was accidentally introduced in<strong>to</strong> the USA and<br />

Europe with seedlings of Asian C. crenata<br />

(Anagnostakis, 1987). Other tree species such as<br />

oak (Quercus spp.) are occasionally attacked,<br />

although only with minor commercial damage.<br />

The fungus is a wound pathogen, colonizing<br />

the bark and cambium tissues as a spreading<br />

mycelium. The host defence reaction produces<br />

cankers (see Fig. 12.18), and branches die if<br />

cankers girdle their circumference. Cankers are<br />

often coloured reddish-brown due <strong>to</strong> the abundant<br />

formation of conidia which ooze out in<br />

sticky tendrils from the pycnidia producing<br />

them. Conidia are spread by animals and rainsplash.<br />

Perithecia are also formed, embedded<br />

in the bark tissue. The fungus is heterothallic,<br />

with a bipolar mating system. Conidia function<br />

as spermatia but can also germinate directly<br />

<strong>to</strong> cause fresh infections. The roots are not<br />

normally affected, and sucker shoots may grow<br />

from intact roots<strong>to</strong>cks or from points proximal<br />

<strong>to</strong> girdling cankers, but these new shoots also<br />

become infected in due course. Following the<br />

establishment of C. parasitica, the American<br />

chestnut tree has become reduced <strong>to</strong> a shrublike<br />

habit, not unlike the way in which the elm<br />

tree in Europe has been affected by Dutch elm<br />

disease (see p. 366). The disease caused by<br />

C. parasitica is known as chestnut blight and<br />

has had a dramatic effect especially in the<br />

eastern United States, where C. dentata once<br />

accounted for 50% of the value of hardwood<br />

timber (Agrios, 2005).<br />

Hypovirulence in Cryphonectria parasitica<br />

Similarly severe outbreaks <strong>to</strong> those in the United<br />

States were noted in Europe, but self-healing<br />

cankers were observed about 15 years after the<br />

first sighting of the disease. Self-healing was<br />

shown by Grente and Sauret (1969) <strong>to</strong> be<br />

associated with the presence of hypovirulent<br />

C. parasitica isolates, i.e. strains with a much<br />

reduced virulence. Moreover, when the hyphae of<br />

a fully virulent colony were allowed <strong>to</strong> fuse with<br />

those of a hypovirulent strain by anas<strong>to</strong>mosis,<br />

the former became hypovirulent, <strong>to</strong>o. Eventually<br />

it was discovered that hypovirulence is associated<br />

with the presence of an unusual fungal<br />

virus consisting of double-stranded RNA surrounded<br />

by membranes but without a protein<br />

coat (Anagnostakis & Day, 1979). This new type<br />

of virus was named Hypovirus (see Dawe & Nuss,<br />

2001). There are now three species of Hypovirus<br />

known from Cryphonectria (Smart et al., 2000),<br />

the best-examined being CHV-1 (Cryphonectria<br />

Hypovirus 1). Cryphonectria is also a reposi<strong>to</strong>ry<br />

for mycoviruses belonging <strong>to</strong> several other<br />

taxonomic groups (Hillman & Suzuki, 2004).<br />

Strains of C. parasitica infected by Hypovirus<br />

have several phenotypic characteristics that<br />

distinguish them from uninfected C. parasitica.<br />

In addition <strong>to</strong> their hypovirulence on chestnut<br />

trees, these include a pale rather than orange<br />

colony colour on agar, reduced production of an<br />

extracellular laccase and of a cell surface hydrophobin<br />

protein due <strong>to</strong> reduced transcription of<br />

their genes, poor asexual sporulation, and lack of<br />

sexual reproduction because of female sterility.<br />

Several genes were thus found <strong>to</strong> be affected by<br />

the presence of the virus. This pleiotropic effect<br />

is thought <strong>to</strong> be due <strong>to</strong> the action of the virus on<br />

various signalling cascades, whereby the stimulation<br />

of the cAMP pathway due <strong>to</strong> an inhibition<br />

of the inhibi<strong>to</strong>ry a subunit (CPG-1) of a large<br />

trimeric G protein has been particularly strongly<br />

implicated (Chen et al., 1996).<br />

The CHV-1 virus genome consists of two open<br />

reading frames which produce al<strong>to</strong>gether three<br />

proteins (Shapira et al., 1991). It is not yet clear<br />

how these act <strong>to</strong> cause the observed symp<strong>to</strong>ms,<br />

except that the p29 protein is responsible for<br />

reducing the pigmentation and asexual sporulation<br />

as well as transcription of the laccase gene.<br />

However, p29 does not seem <strong>to</strong> cause hypovirulence<br />

(Nuss, 1996). In contrast, when the<br />

CPG-1 protein levels were reduced by the<br />

presence of the wild-type virus or by genetic<br />

manipulation of uninfected C. parasitica, or when


MAGNAPORTHACEAE<br />

377<br />

cAMP levels were raised by chemical treatment,<br />

much of the phenotype of hypovirus infection<br />

including hypovirulence was induced. This<br />

hints at an interplay between several signalling<br />

pathways, as outlined for Magnaporthe grisea<br />

(Fig. 12.48).<br />

Experiments with CHV-1 have also provided<br />

an insight in<strong>to</strong> the sexual reproductive system<br />

of C. parasitica because the virus suppresses the<br />

genes encoding the mating pheromones which<br />

could thus be identified (Zhang et al., 1998). Both<br />

male and female pheromones were suppressed<br />

by CHV-1 infections. The reason why infection<br />

leads selectively <strong>to</strong> female sterility is unclear<br />

but may be because a certain proportion of<br />

conidia, which double up as spermatia, remains<br />

uninfected by the virus during conidiogenesis.<br />

Biological control of Cryphonectria parasitica<br />

by Hypovirus<br />

The transmission of Hypovirus from one strain of<br />

C. parasitica <strong>to</strong> another is mediated by anas<strong>to</strong>mosis,<br />

which requires vegetative compatibility.<br />

In C. parasitica as in Podospora anserina (p. 320),<br />

there are several genetic loci controlling vegetative<br />

compatibility, and anas<strong>to</strong>mosis as well<br />

as virus transmission occur readily between two<br />

strains possessing identical alleles at all six vegetative<br />

incompatibility (vic) loci. Heteroallelism<br />

at one or more loci restricts anas<strong>to</strong>mosis and<br />

reduces the percentage of virus transmission,<br />

whereby mismatches at certain loci have a more<br />

restrictive effect than those at others (Cortesi<br />

et al., 2001). Of course, virus transmission in the<br />

field will be higher in populations containing<br />

a low diversity of vic alleles. Such is the case<br />

in Europe, where hypovirulent strains were<br />

observed <strong>to</strong> spread rapidly after their discovery.<br />

There are different strains of Hypovirus which<br />

may be mildly hypovirulent, i.e. only moderately<br />

restricting canker development and asexual<br />

sporulation in Cryphonectria, or may be aggressive.<br />

The latter permit Cryphonectria <strong>to</strong> form<br />

only very small cankers, but also greatly reduce<br />

asexual sporulation. Since the virus can be disseminated<br />

at least <strong>to</strong> a certain extent in conidia<br />

of C. parasitica, mild virus strains may spread<br />

faster in nature than aggressive ones.<br />

In addition <strong>to</strong> relying on the natural spread<br />

of hypovirulent strains of C. parasitica in Europe,<br />

it has proven possible <strong>to</strong> implement a biological<br />

control strategy by inoculating active cankers<br />

with a paste containing a mixture of hypovirulent<br />

strains differing in their vic alleles.<br />

Success has been obtained especially in chestnut<br />

orchards or in regions where hypovirulent<br />

strains were rare in the field (Heininger &<br />

Rigling, 1994). Active cankers can be converted<br />

in<strong>to</strong> healing cankers if one of the inoculated<br />

hypovirulent strains can anas<strong>to</strong>mose with the<br />

fully pathogenic strain. In addition, the natural<br />

spread of hypovirulence was observed around<br />

sites of release (Heininger & Rigling, 1994).<br />

In contrast, in the eastern USA where a<br />

great diversity of vic alleles exists in the wild,<br />

biological control measures have not generally<br />

been successful except in isolated forests.<br />

Another reason for the difficulties may be that<br />

an aggressive hypovirus strain was chosen for<br />

initial release experiments, which strongly<br />

reduced the ability of C. parasitica <strong>to</strong> produce<br />

conidia (Nuss, 1992). Current strategies are using<br />

the fact that cDNA of Hypovirus can be stably<br />

integrated in<strong>to</strong> the genome of C. parasitica,<br />

and produces double-stranded viral RNA in the<br />

fungus. Whereas the RNA of the virus is not<br />

transmissible via sexual reproduction in<strong>to</strong> ascospores,<br />

the integrated genomic DNA, of course,<br />

is transmitted so long as the viral RNA, which<br />

causes female sterility, is absent from the fungal<br />

cy<strong>to</strong>plasm. This strategy promises <strong>to</strong> be more<br />

successful than previous release experiments<br />

because the vic alleles are re-mixed during<br />

sexual reproduction, thereby facilitating the<br />

introduction of the virus in<strong>to</strong> a population<br />

with diverse vic alleles (Dawe & Nuss, 2001).<br />

12.9 Magnaporthaceae<br />

This small family (9 genera, 26 species) is<br />

currently homeless, having been excluded from<br />

the Diaporthales (see p. 373) with which it<br />

was formerly thought <strong>to</strong> be associated (Berbee,<br />

2001; Castlebury et al., 2002). We include it here


378 HYMENOASCOMYCETES: PYRENOMYCETES<br />

because two members, Magnaporthe grisea and<br />

Gaeumannomyces graminis, are important plant<br />

pathogens. Seminal work on developmental and<br />

molecular aspects of plant pathogenesis has<br />

been done with M. grisea which will be the<br />

main focus of this section.<br />

Magnaporthe grisea causes rice blast disease<br />

in which individual infections give rise <strong>to</strong><br />

spindle-shaped necrotic lesions on rice leaves.<br />

The fungus is particularly common in South<br />

East Asia, its probable centre of origin where<br />

it has been known for centuries (Rao, 1994). Rice<br />

blast has spread <strong>to</strong> virtually all rice-growing<br />

areas, although it is more severe in cooler<br />

climates. Considerable research efforts are<br />

being directed at controlling rice blast, although<br />

the output in the shape of fungicides against<br />

M. grisea or the seeds of resistant rice cultivars<br />

may well be beyond the financial means of the<br />

small-scale agricultural systems found in many<br />

countries in Asia and Africa where rice blast is<br />

a problem.<br />

In the field, the fungus is encountered mainly<br />

in the anamorph state which used <strong>to</strong> be called<br />

Pyricularia oryzae if growing on rice. In addition<br />

<strong>to</strong> rice, M. grisea can attack wheat, barley and<br />

various wild grasses on which the asexual state is<br />

called P. grisea. Accordingly, the rice pathogen<br />

should perhaps be renamed M. oryzae, but we feel<br />

bound by convention <strong>to</strong> retain M. grisea since<br />

that name is universally used. There are several<br />

strains with different host spectra, and e.g. the<br />

wheat blast strain in Brazil is genetically distinct<br />

from rice blast strains present in the same<br />

regions (Urashima et al., 1993). This is consistent<br />

with the suggestion by Couch et al. (2005)<br />

that the rice-infecting lineage of M. grisea arose<br />

from a single host-switching event from Setaria<br />

millet early in the his<strong>to</strong>ry of rice cultivation<br />

which began around 5000 BC. Sexual reproduction<br />

is rare in the field especially with the rice<br />

strains, so that discrete clones of M. grisea<br />

populations are typically formed in many ricegrowing<br />

areas (Zeigler, 1998). In tropical climates<br />

with up <strong>to</strong> three cropping seasons each year, the<br />

fungus can continuously infect fresh green<br />

foliage, whereas in Southern Europe it overwinters<br />

on rice stubble. Magnaporthe grisea is<br />

haploid, with each nucleus containing about<br />

seven chromosomes (Valent, 1997).<br />

12.9.1 Conidium germination and<br />

appressorium formation in<br />

Magnaporthe grisea<br />

A summary of the disease cycle is given in<br />

Fig. 12.44. Good reviews have been written by<br />

Howard and Valent (1996), Valent (1997) and<br />

Tucker and Talbot (2001). The conidia of M. grisea<br />

are three-celled, with each cell containing a<br />

single nucleus. A mature conidium swells upon<br />

hydration, and this causes the breakage of the<br />

wall at the tip of the spore, releasing a drop<br />

of mucilage s<strong>to</strong>red in the periplasmic space<br />

(Fig. 12.45a). This may already occur while the<br />

spore is still attached <strong>to</strong> the conidiophore.<br />

The exact chemical nature of this mucilage is<br />

unknown, although it probably contains glycoproteins.<br />

It attaches the spore firmly <strong>to</strong> the wax<br />

of the host cuticle or other hydrophobic surfaces<br />

(Hamer et al., 1988; Howard, 1994). Mucilage<br />

release is a purely physical phenomenon which<br />

does not require any de novo metabolic<br />

activities.<br />

Under suitable conditions, a single conidial<br />

cell usually the basal or apical cell, more rarely<br />

the central one emits a germ tube. Numerous<br />

conflicting reports have been published on the<br />

requirements for germination, but in our experience<br />

conidia germinate readily in aqueous<br />

suspension or in contact with any inert surface,<br />

provided that they have been washed by centrifugation<br />

and resuspension in water. Washing<br />

removes an au<strong>to</strong>-inhibi<strong>to</strong>r which prevents germination<br />

of spores in dense suspensions (Kono et al.,<br />

1991). Germination on the plant cuticle may<br />

appear <strong>to</strong> be stimulated simply because the inhibi<strong>to</strong>r<br />

is lipophilic and dissolves in<strong>to</strong> the cuticular<br />

waxes, thereby becoming diluted from the<br />

spore (Hedge & Kolattukudy, 1997).<br />

In contrast <strong>to</strong> spore germination, commitment<br />

<strong>to</strong> appressorium formation requires<br />

the presence of specific environmental signals.<br />

These are perceived after the germ tube has<br />

formed a hook-like appressorium initial (Bourett<br />

& Howard, 1990; deZwaan et al., 1999).<br />

Appressoria are formed from hooks upon contact


MAGNAPORTHACEAE<br />

379<br />

Fig12.44 The infection sequence of<br />

Magnaporthe grisea.The approximate<br />

times are indicated, as are developmental<br />

stages at which signalling cascades are<br />

known <strong>to</strong> be involved.


380 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.45 Microscopy of Magnaporthe grisea. (a) Hydrated conidium of M. grisea showing the drop of mucilage exuded at the apex.<br />

(b) Ungerminated conidium with dense cy<strong>to</strong>plasmic contents and developing vacuoles. (c) Conidium on a hydrophobic surface about<br />

18 h after germination. Most cy<strong>to</strong>plasmic contents have been translocated in<strong>to</strong> the appressorium which is laying down a dark<br />

melanized wall. (d) SEM of a mature appressorium of M. grisea.Whereas the conidial cells have collapsed due <strong>to</strong> the loss of turgor<br />

pressure, the appressorium is still turgid. (e) TEM of an appressorium penetrating a cellophane membrane.The appressorial wall is<br />

heavily melanized with the exception of the basal region through which the penetration peg has emerged. A large vacuole (Vac) and<br />

nucleus (Nuc) are visible inside the appressorium. (f,g) Formation of bulbous hyphae from an appressorium, some 36 h after<br />

penetration of an onion epidermis. A thinner secondary hypha has already been produced (arrow). Micrographs in (d) and (e) kindly<br />

provided by R. J. Howard; (d) reprinted fromValent (1997) with kind permission of Springer Science and Business Media;<br />

(e) reprinted from Bourett and Howard (1990), by copyright permission of the National Research Council of Canada.Original<br />

micrographs in f and g kindly supplied by A. J.Foster. (a c) <strong>to</strong> same scale; (f,g) <strong>to</strong> same scale.


MAGNAPORTHACEAE<br />

381<br />

with a hydrophobic surface, or by means of<br />

chemical cues such as cutin monomers (especially<br />

1,16-hexadecanediol). Major carbon and<br />

energy s<strong>to</strong>rage products in spores of M. grisea are<br />

cy<strong>to</strong>plasmic glycogen deposits and lipid droplets.<br />

During and after germ tube emergence, the<br />

glycogen becomes hydrolysed and enters the<br />

germ tube as soluble sugars. In contrast, lipid<br />

droplets become mobilized and migrate intact<br />

in<strong>to</strong> the germ tube (Thines et al., 2000), most<br />

probably along elements of the cy<strong>to</strong>skele<strong>to</strong>n.<br />

Lipid droplets accumulate in the developing<br />

appressorium (Figs. 12.45b,c), and their exit<br />

is prevented by the formation of a septum<br />

which cuts off the appressorial cy<strong>to</strong>plasm from<br />

that of the germ tube. The single nucleus present<br />

in the germinating conidial cell divides<br />

once, and one nucleus enters the appressorium<br />

whereas the other remains in the germ tube.<br />

After the septum has been completed, the germ<br />

tube and conidial cell collapse (Fig. 12.45d).<br />

12.9.2 Appressorium maturation<br />

and host penetration by<br />

Magnaporthe grisea<br />

The maturation of the appressorium is a rapid<br />

process, and within 12 h of conidia being placed<br />

on a suitable surface, mature appressoria will<br />

have formed (Fig. 12.44). Penetration takes place<br />

about 24 h after spore germination. Several<br />

processes can be observed with the light microscope,<br />

e.g. when drops of a conidial suspension<br />

are placed on plastic coverslips or onion epidermis.<br />

The lipid droplets which have accumulated<br />

in the appressorium aggregate in<strong>to</strong> larger drops.<br />

In the centre of the appressorium, a vacuole<br />

forms and enlarges, and the lipid droplets enter<br />

the vacuole by microau<strong>to</strong>phagocy<strong>to</strong>sis and are<br />

degraded there (Weber et al., 2001). This presumably<br />

provides the energy needed for the penetration<br />

events which follow. While the lipid<br />

droplets are being degraded, a thick brown<br />

wall forms on the outside of the initially<br />

hyaline wall. This contains melanin. The melanin<br />

of M. grisea, like that of most ascomycetes, is<br />

a polymer of dihydroxynaphthalene (DHN),<br />

although other pathways exist in other groups<br />

of fungi, plants and animals. DHN is a<br />

pentaketide, i.e. it is formed by the head-<strong>to</strong>-tail<br />

condensation of five acetate units (Fig. 12.46)<br />

which may well arise directly from lipid oxidation.<br />

Melanin biosynthesis in fungi has been<br />

reviewed by Bell and Wheeler (1986) and Butler<br />

and Day (1998). In addition <strong>to</strong> giving strength<br />

<strong>to</strong> the appressorial wall, melanin also reduces<br />

the pore size <strong>to</strong> less than 1 nm so that molecules<br />

larger than water cannot traverse the<br />

melanized wall.<br />

Soon after the melanin layer has been<br />

deposited, the appressorium begins <strong>to</strong> synthesize<br />

large quantities of solutes, especially glycerol<br />

(de Jong et al., 1997). Since these solutes cannot<br />

escape, water moves inwards by osmosis, resulting<br />

in the generation of a turgor pressure of<br />

up <strong>to</strong> 8 MPa (¼ 80 bar). This is one of the highest<br />

pressures recorded in any living cell (Howard<br />

et al., 1991). At its base, the appressorium is<br />

cemented tightly <strong>to</strong> the surface of the cuticle by<br />

a very effective adhesive which consists mainly<br />

of glycoproteins and lipids (Ohtake et al., 1999).<br />

In the basal plate of the appressorium, there is<br />

a small non-melanized pore, so that eventually<br />

the turgor pressure is relieved by driving a penetration<br />

peg through the surface (Fig. 12.45e).<br />

The penetration peg contains a conspicuous<br />

internal skele<strong>to</strong>n of actin (Bourett & Howard,<br />

1992), which is almost certainly involved in<br />

maintaining its shape, akin <strong>to</strong> the internal<br />

skele<strong>to</strong>n at the tip of a growing hypha (see<br />

Fig. 1.8).<br />

12.9.3 Colonization of the host<br />

and sporulation of<br />

Magnaporthe grisea<br />

Once the cuticle has been penetrated, the<br />

infection peg enlarges and branches <strong>to</strong> form<br />

hyphae which initially look swollen and are<br />

therefore called bulbous hyphae. Interestingly,<br />

at this stage the ultrastructure of an infection by<br />

M. grisea is similar <strong>to</strong> that of haus<strong>to</strong>ria produced<br />

by biotrophic pathogens such as rusts or powdery<br />

mildews in that the bulbous hyphae do not<br />

pierce but invaginate the host plasmalemma<br />

(Heath et al., 1992). This initial biotrophic phase<br />

is confined <strong>to</strong> the first epidermal cell encountered;<br />

hyphae reaching adjacent cells quickly


382 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.46 The biosynthetic pathway<br />

of fungal melanin via DHN<br />

(dihydroxynaphthalene). A polyketide<br />

synthase fuses five acetate units<br />

head-<strong>to</strong>-tail followed by cyclization,<br />

resulting in the pentaketide<br />

1,3,6,8 -tetrahydroxynaphthalene.<br />

This is modified by two reductases<br />

and two dehydratases <strong>to</strong><br />

1,8 -dihydroxynaphthalene.This melanin<br />

precursor, which is presumably<br />

synthesized in the cy<strong>to</strong>plasm, is<br />

released in<strong>to</strong> the wall and then<br />

polymerized <strong>to</strong> melanin by wall-bound<br />

oxidases, presumably laccases.<br />

Melanin biosynthesis is an excellent<br />

target for fungicides, and the points<br />

of inhibition of several of them are<br />

indicated. Based partly on Uesugi (1998)<br />

and Thompson et al. (2000).


MAGNAPORTHACEAE<br />

383<br />

Fig12.47 (a) Conidiophores of<br />

Magnaporthe grisea emerging from<br />

a s<strong>to</strong>ma on a rice leaf. Detached<br />

conidia are also shown (b).<br />

colonize the plant tissue. These secondary colonizing<br />

hyphae are thinner and straighter than<br />

the bulbous hyphae (Figs. 12.45f,g). Within 5 days<br />

of infection, diamond-shaped lesions of dead<br />

tissue develop, and these emit conidiophores<br />

which often grow out through dead s<strong>to</strong>mata<br />

(Fig. 12.47).<br />

The interaction between M. grisea and rice<br />

is governed by a gene-for-gene relationship<br />

(Valent, 1997; see p. 619). In compatible interactions,<br />

the host response is delayed because the<br />

fungus is not immediately recognized by the<br />

plant. However, even virulent strains of M. grisea<br />

will eventually be exposed <strong>to</strong> the <strong>to</strong>xic products<br />

(phy<strong>to</strong>alexins) of the host’s delayed immune<br />

response, and M. grisea seems <strong>to</strong> have evolved<br />

mechanisms <strong>to</strong> deal with phy<strong>to</strong>alexins by exclusion.<br />

It is becoming clear that ABC transporters<br />

play a crucial role, and we have already come<br />

across them as a resistance mechanism developed<br />

by Candida albicans against clinical drugs<br />

(see p. 278). The report by Urban et al. (1999)<br />

on M. grisea was one of the first <strong>to</strong> implicate<br />

a phy<strong>to</strong>alexin-excluding ABC transporter as an<br />

important fac<strong>to</strong>r in the colonization of a plant<br />

host by a fungal pathogen. ABC transporters<br />

have now been shown <strong>to</strong> play a crucial role in the<br />

exclusion of <strong>to</strong>xic substances as well as fungicides<br />

in several plant-pathogenic fungi (Hayashi<br />

et al., 2002; Stergiopoulos et al., 2002).


384 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.48 The three principal signalling cascades involved in appressorium differentiation in Magnaporthe grisea and in many other<br />

differentiation processes in other eukaryotes.Cross-talk occurs between the individual pathways, but this is not shown here.<br />

Abbreviations are as follows: CaM (calmodulin), cAMP (cyclic adenosine monophosphate),Cat. (catalytic), DAG (diacylglycerol),<br />

GTP (guanosine triphosphate), IP 3<br />

(inosi<strong>to</strong>l trisphosphate), MAPK (mi<strong>to</strong>gen-activated protein kinase), PIP 2<br />

(phosphatidylinosi<strong>to</strong>l<br />

bisphosphate), Reg. (regula<strong>to</strong>ry), TF (transcription fac<strong>to</strong>r).<br />

12.9.4 Signalling and pathogenesis<br />

in Magnaporthe grisea<br />

Both signals for appressorium initiation surface<br />

hydrophobicity and cutin monomers are<br />

probably perceived at the plasma membrane of<br />

the Magnaporthe germ tube. The actual recep<strong>to</strong>rs<br />

are unknown at present, but the Pth11p plasma<br />

membrane protein is likely <strong>to</strong> be one of them<br />

(deZwaan et al., 1999). The transmission of signals<br />

from the plasma membrane <strong>to</strong> the nucleus<br />

occurs along several different routes (Fig. 12.48;<br />

Dean, 1997; Tucker & Talbot, 2001) which are<br />

briefly outlined below:<br />

1. One membrane recep<strong>to</strong>r receiving the<br />

chemical stimulus 1,16-hexadecanediol acts via<br />

a trimeric GTP-binding protein <strong>to</strong> activate an<br />

adenylate cyclase which converts ATP in<strong>to</strong> the<br />

second messenger, cyclic AMP (cAMP). This, in<br />

turn, activates a protein kinase A by releasing<br />

its monomeric catalytic subunits from the<br />

inactive tetramer, and the catalytic subunits<br />

then phosphorylate regula<strong>to</strong>ry proteins (transcription<br />

fac<strong>to</strong>rs) which enter the nucleus<br />

and activate the genes required for specific<br />

developmental steps.<br />

2. The involvement of a second common<br />

eukaryotic signalling cascade in appressorium<br />

formation was suggested by Thines et al. (1997)<br />

who noted that diacylglycerols could trigger<br />

appressorium formation on normally non-inductive<br />

hydrophilic surfaces such as glass. The<br />

initial signal (hydrophobicity) is thus likely <strong>to</strong><br />

be transduced via phospholipase C which hydrolyses<br />

the membrane lipid phosphatidylinosi<strong>to</strong>l<br />

in<strong>to</strong> inosi<strong>to</strong>l-triphosphate (IP 3 ) and diacylglycerol<br />

(DAG). IP 3 acts by releasing Ca 2þ from


MAGNAPORTHACEAE<br />

385<br />

intracellular s<strong>to</strong>res such as the endoplasmic<br />

reticulum; this forms a complex with the<br />

calcium-binding protein calmodulin, and the<br />

calcium calmodulin complex activates a calmodulin-dependent<br />

protein kinase. The DAG component<br />

directly activates protein kinase C (Fig.<br />

12.48). Both the IP 3 and DAG branches of this<br />

signalling pathway thereby lead <strong>to</strong> the activation<br />

of protein kinases which phosphorylate transcription<br />

fac<strong>to</strong>rs.<br />

3. Mi<strong>to</strong>gen-activated protein kinase (MAPK)<br />

pathways have been proposed <strong>to</strong> be involved at<br />

various points in appressorium induction and<br />

maturation in M. grisea (Xu & Hamer, 1996;<br />

Thines et al., 2000). A mi<strong>to</strong>gen is an extracellular<br />

substance which stimulates nuclear division<br />

or cell differentiation; here it is a stimulus<br />

for appressorium formation. The MAP kinase<br />

encoded by PMK1 responds <strong>to</strong> a surface signal and<br />

interacts with the cAMP pathway in a manner<br />

not yet entirely unders<strong>to</strong>od, <strong>to</strong> initiate appressorium<br />

formation (see Fig. 12.48). A second MAP<br />

kinase, MPS1, is involved in penetration of the<br />

epidermis from mature appressoria. A third MAP<br />

kinase, OSM1, is involved in turgor regulation<br />

during osmotic stress but plays no role in<br />

appressorium functioning (Dixon et al., 1999).<br />

MAP kinases are very highly conserved between<br />

different fungi, <strong>to</strong> the extent that they are<br />

functional when their genes are exchanged,<br />

e.g. between M. grisea, Candida albicans and<br />

Saccharomyces cerevisiae (Xu, 2000).<br />

It should be noted that any one of the above<br />

principal signalling cascades may act repeatedly<br />

in the course of appressorium development, as<br />

indicated in Fig. 12.44, and in other events such<br />

as production of conidiophores and conidia.<br />

Signal cascades acting repeatedly in the life<br />

cycle of M. grisea use many shared components<br />

and only a few specific ones at any one time<br />

point. There is also considerable cross-talk<br />

between the three types of signalling cascade<br />

mentioned here (see Kronstad et al., 1998), and<br />

equivalent signalling pathways are involved<br />

in infection processes of many other fungal<br />

pathogens and in other fundamental processes<br />

such as yeast hyphal dimorphism, mating and<br />

osmoregulation in most fungi examined <strong>to</strong><br />

date (Xu, 2000). They are also fundamentally<br />

conserved across other eukaryotic life forms,<br />

and some are even found in prokaryotes. For<br />

this reason, signalling cascades are unlikely <strong>to</strong><br />

provide suitably specific targets for fungicides,<br />

and interest in this aspect of signalling seems <strong>to</strong><br />

have waned somewhat in recent years.<br />

12.9.5 Gaeumannomyces graminis<br />

Gaeumannomyces graminis var. tritici causes take-all<br />

disease of wheat, barley, rye and numerous<br />

wild grasses. The pathogen is soil-borne and<br />

Fig12.49 Melanized hyphopodia of<br />

Gaeumannomycesgraminis arising from an<br />

equally strongly melanized runner hypha<br />

formed on the hydrophobic surface of a<br />

plastic coverslip.


386 HYMENOASCOMYCETES: PYRENOMYCETES<br />

Fig12.50 Diagrammatic representation of the<br />

relationship between wheat yield (solid line), the<br />

severity of take all infection (dashed line) and the<br />

occurrence of 2,4-diacetylphloroglucinol-producing<br />

Pseudomonas spp. (dotted line) on a field with successive<br />

wheat cultivation.The insert shows the molecular<br />

structure of 2,4-diacetylphloroglucinol. Based partly<br />

on Parry (1990) and Weller et al.(2002).<br />

infects the roots, causing blackening and decay.<br />

As a result, the cereal plants appear stunted and<br />

produce small, non-fertile heads which appear<br />

bleached and are therefore called ‘white heads’.<br />

The fungus overwinters on cereal stubble and<br />

infects the new crop mainly in spring. The infection<br />

mechanism is not entirely clear, and the<br />

fungus may enter host roots directly from dark<br />

melanized hyphae (runner hyphae) or melanized<br />

hyphopodia which may be aggregated <strong>to</strong> form<br />

infection cushions or mycelial mats (Butler &<br />

Jones, 1949). A hyphopodium is defined as an<br />

appressorium produced from a vegetative hypha<br />

rather than a germinating spore (Fig. 12.49). The<br />

turgor pressure in hyphopodia is around 1.5 MPa<br />

(¼ 15 bar), i.e. considerably less than in appressoria<br />

of M. grisea (Money et al., 1998).<br />

Like M. grisea, G. graminis is a complex of<br />

strains with different but overlapping host<br />

ranges, and varieties tritici, avenae, graminis and<br />

maydis are distinguished. The conidial state of<br />

Gaeumannomyces is a Phialophora, but since the<br />

assignments of anamorphs is difficult, it may be<br />

best <strong>to</strong> speak of the Gaeumannomyces Phialophora<br />

complex (Bryan et al., 1995). Gaeumannomyces<br />

is closely related <strong>to</strong> Magnaporthe (Bryan et al.,<br />

1995).<br />

Take-all is considered <strong>to</strong> be the most important<br />

cereal disease in temperate climates. If<br />

wheat is grown on a field for about four successive<br />

years, the disease will build up <strong>to</strong> very<br />

high levels, causing crop losses in excess of 50%<br />

(Polley & Clarkson, 1980), but if cultivation is<br />

continued the disease will decline <strong>to</strong> an acceptable<br />

base level (Fig. 12.50). If a crop rotation<br />

is carried out, the antagonistic properties of<br />

the soil are lost. The reasons for this remarkable<br />

phenomenon are now beginning <strong>to</strong> be unders<strong>to</strong>od,<br />

and it is clear that the establishment of an<br />

antagonistic soil microflora plays an important<br />

role. In particular, fluorescent Pseudomonas<br />

spp., i.e. pseudomonads which produce watersoluble<br />

substances which fluoresce green<br />

or yellow, have been implicated as agents of<br />

take-all decline (Weller et al., 2002). The metabolite<br />

2,4-diacetylphloroglucinol (Fig. 12.50)<br />

produced by them seems <strong>to</strong> be responsible for<br />

suppression of G. graminis in the rhizosphere<br />

(Raajmakers & Weller, 1998). Wheat roots injured<br />

by Gaeumannomyces are colonized extensively<br />

by Pseudomonas spp., thereby explaining why a<br />

severe take-all attack must occur in order <strong>to</strong><br />

render the soil suppressive.<br />

12.10 Glomerellaceae<br />

Like the Magnaporthaceae, the family<br />

Glomerellaceae is a group of plant-pathogenic<br />

fungi which cannot be assigned <strong>to</strong> any order at<br />

present, although it is certain that it belongs <strong>to</strong><br />

the core group of Pyrenomycetes like all other<br />

groups described in this chapter (Wanderlei-Silva<br />

et al., 2003). The family comprises only one


GLOMERELLACEAE<br />

387<br />

Fig12.51 Colle<strong>to</strong>trichum graminicola. (a) Portion of Sorghum<br />

grain bearing acervuli. (b) An acervulus. (c) Phialides and<br />

phialoconidia.<br />

teleomorph genus, Glomerella, which consists of<br />

five species, and numerous species belonging<br />

<strong>to</strong> the form-genus Colle<strong>to</strong>trichum. Early workers<br />

defined Colle<strong>to</strong>trichum spp. mainly by their association<br />

with host plants, so that approximately<br />

750 ‘species’ were described before von Arx (1957)<br />

tidied them up by synonymizing them in<strong>to</strong><br />

11 taxa. For C. gloeosporioides alone, von Arx<br />

(1957) recognized 600 synonyms. A more recent<br />

treatment is that by Sut<strong>to</strong>n (1992) in which<br />

39 species are described. Many of these are<br />

probably species complexes, and for example<br />

six formae speciales have been described for<br />

C. gloeosporioides.<br />

Glomerella forms dark-walled perithecia which<br />

occur singly. They are long-necked and release<br />

ascospores passively. The Colle<strong>to</strong>trichum state is an<br />

acervulus, i.e. a saucer-shaped conidioma which<br />

develops within the host tissue, rupturing<br />

the cuticle at maturity (Fig. 12.51a). The<br />

conidiogenous cells form a closely packed palisade<br />

of phialides (Fig. 12.51c). The curved<br />

elongated phialoconidia are produced in a<br />

slimy droplet which is held in place by stiff<br />

dark setae surrounding the acervulus (Fig.<br />

12.51b). This mucilage contains the au<strong>to</strong>inhibi<strong>to</strong>r<br />

mycosporine-alanine which prevents the<br />

germination of conidia until it has been diluted<br />

out (Leite & Nicholson, 1992). Other germination<br />

au<strong>to</strong>inhibi<strong>to</strong>rs are localized within the conidial<br />

cy<strong>to</strong>plasm of diverse Colle<strong>to</strong>trichum spp. (García-<br />

Pajón & Collado, 2003). The presence of such<br />

substances explains why the spores of many<br />

fungi germinate better after they have been<br />

washed.<br />

Mating systems are complex and variable in<br />

Glomerella. For example, in G. cingulata (anamorph<br />

C. gloeosporioides), both homothallic and heterothallic<br />

strains occur, and there may be more<br />

than two mating type idiomorphs (Cisar &<br />

TeBeest, 1999). The occurrence of multiple alleles<br />

at one mating type locus is common in the<br />

Basidiomycota but has apparently not been<br />

found in any ascomycete other than Glomerella.<br />

Many strains belonging <strong>to</strong> a given species


388 HYMENOASCOMYCETES: PYRENOMYCETES<br />

complex appear <strong>to</strong> be reproductively isolated; for<br />

instance, the sexual state of C. lindemuthianum<br />

(teleomorph G. cingulata f. sp. phaseoli) has not<br />

been found in nature but can be readily induced<br />

in agar culture by the pairing of compatible<br />

isolates (Roca et al., 2003).<br />

Species of Colle<strong>to</strong>trichum cause serious diseases<br />

on a wide range of plants. These are often referred<br />

<strong>to</strong> as anthracnose because of the appearance<br />

of sunken necrotic lesions, as exemplified by<br />

C. lindemuthianum which causes anthracnose of<br />

beans, peas and other legumes (Plate 5i). These<br />

necrotic lesions contain the acervuli. Other<br />

important pathogenic species are C. gloeosporioides<br />

(anthracnose of a range of tropical fruits and<br />

many other plants), C. coffeanum (coffee berry<br />

disease), C. gossypii (boll rot and anthracnose of<br />

cot<strong>to</strong>n), C. musae (post-harvest fruit anthracnose<br />

on banana), C. graminicola (anthracnose of maize<br />

and sorghum), and C. coccodes (anthracnose of<br />

<strong>to</strong>ma<strong>to</strong>, black dot disease of pota<strong>to</strong>). Many of<br />

these produce phy<strong>to</strong><strong>to</strong>xic substances which are<br />

involved in causing disease symp<strong>to</strong>ms (García-<br />

Pajón & Collado, 2003). Whilst diseases in the<br />

field are important, post-harvest rots probably<br />

cause even greater economic damage, especially<br />

in the tropics. This comes about because<br />

Colle<strong>to</strong>trichum spp. can cause latent infections<br />

which give rise <strong>to</strong> disease symp<strong>to</strong>ms only during<br />

fruit ripening in s<strong>to</strong>rage. Good descriptions of<br />

anthracnose diseases may be found in two<br />

volumes dedicated <strong>to</strong> Colle<strong>to</strong>trichum, which is<br />

one of the most important genera of fungal<br />

plant pathogens worldwide (Bailey & Jeger, 1992;<br />

Prusky et al., 2000). Several Colle<strong>to</strong>trichum spp. have<br />

been developed with limited success as biocontrol<br />

agents against weeds (Watson et al., 2000).<br />

12.10.1 Infection strategies in Colle<strong>to</strong>trichum<br />

Most species infect their host from germinating<br />

conidia which emit germ tubes terminating in a<br />

melanized appressorium. The details of appressorium<br />

formation and function are very similar<br />

<strong>to</strong> those in Magnaporthe grisea (see pp. 378 381),<br />

as far as they are known (Bailey et al., 1992).<br />

One of the fascinations of the genus<br />

Colle<strong>to</strong>trichum lies in the range of postappressorial<br />

infection strategies which various<br />

species have evolved. These have been well<br />

described by O’Connell et al. (2000) and<br />

Latunde-Dada (2001), and are briefly summarized<br />

below.<br />

Necrotrophic pathogens<br />

Colle<strong>to</strong>trichum capsici is a necrotrophic pathogen<br />

infecting red peppers. It shows intercellular<br />

growth which commences immediately after<br />

penetration and is accompanied by the secretion<br />

of cell wall-degrading enzymes (Pring et al.,<br />

1995).<br />

Hemibiotrophic pathogens<br />

The pattern of infection by numerous species<br />

(exemplified by C. lindemuthianum) is very similar<br />

<strong>to</strong> that in Magnaporthe grisea (p. 381). The penetration<br />

peg arising from an appressorium forms<br />

a vesicle which emits swollen primary hyphae<br />

inside the first-colonized epidermal cell and in<br />

adjacent cells (see Figs. 12.45f,g). The primary<br />

hyphae do not breach the plant plasma membrane<br />

but invaginate it, and there is evidence of<br />

a distinct matrix between the fungal hypha and<br />

the host membrane, in analogy <strong>to</strong> the haus<strong>to</strong>rium<br />

formed by biotrophic fungal pathogens<br />

(O’Connell et al., 1985; Mendgen & Hahn, 2002).<br />

After 1 3 days, the biotrophic phase breaks<br />

down and thinner secondary hyphae are<br />

formed which spread the infection, forming<br />

a necrotrophic lesion. Secondary hyphae differ<br />

fundamentally in surface properties from primary<br />

hyphae and are not surrounded by an<br />

extracellular matrix (Perfect et al., 2001). A<br />

modification of this pattern is observed, e.g. in<br />

C. destructivum on cowpea (Vigna unguiculata),<br />

in which the biotrophic stage is confined <strong>to</strong><br />

one epidermal cell containing a multi-lobed<br />

vesicle, from which secondary hyphae initiate<br />

the necrotrophic phase (Latunde-Dada et al.,<br />

1996).<br />

Post-harvest pathogens<br />

This group of Colle<strong>to</strong>trichum spp. is responsible for<br />

most of the diseases of ripe fruits, especially in<br />

tropical areas. An example is C. musae on banana<br />

fruits. Spores alighting on unripe fruits prior<br />

<strong>to</strong> harvest may germinate and form appressoria<br />

and even penetration pegs, but there the


GLOMERELLACEAE<br />

389<br />

infection process stalls. Infection is continued on<br />

the ripened fruits after a period of s<strong>to</strong>rage.<br />

This infection strategy enables the pathogen <strong>to</strong><br />

avoid the high levels of phy<strong>to</strong>alexins in unripe<br />

fruits (Latunde-Dada, 2001). Resumption of development<br />

is triggered by the strong increase in<br />

ethylene levels associated with fruit ripening<br />

(Flaishman & Kolattukudy, 1994).<br />

Endophytic colonization<br />

Several anthracnose pathogens including<br />

C. musae, C. coccodes and strains of C. gloeosporioides<br />

develop a prolonged latent phase inside their<br />

host plant (Rodriguez & Redman, 1997).<br />

Colonization during the endophytic phase may<br />

begin by infection through s<strong>to</strong>mata rather than<br />

direct penetration of the epidermis, and intercellular<br />

colonization of host tissue. During<br />

senescence of the host plant, the host’s<br />

immune system degenerates and active colonization<br />

may ensue, resulting in the development of<br />

sporulating anthracnose-type lesions. Freeman<br />

and Rodriguez (1993) and Redman et al. (1999a)<br />

have shown that the deletion of a single gene<br />

can convert an aggressive pathogen (C. magna)<br />

<strong>to</strong> a permanently symp<strong>to</strong>mless endophyte.<br />

This finding emphasizes the possibility that<br />

‘endophytic mutualism is only a gene away<br />

from pathogenicity’ (Latunde-Dada, 2001), and<br />

also indicates one mechanism by which endophytes<br />

may evolve in nature. A further fundamental<br />

point of interest is that colonization by<br />

such symp<strong>to</strong>mless endophytes renders the host<br />

plant resistant against infection by pathogenic<br />

strains of the same species, and also other fungal<br />

pathogens (Redman et al., 1999b).


13<br />

Hymenoascomycetes: Erysiphales<br />

13.1 <strong>Introduction</strong><br />

The Erysiphales are a clearly defined, monophyletic<br />

order of about 500 species, all of which are<br />

obligately biotrophic pathogens of plants. Braun<br />

(1987, 1995) has provided thorough monographic<br />

treatments of this group, including descriptions<br />

of most known species. There is only one family,<br />

the Erysiphaceae. The species grouped here cause<br />

symp<strong>to</strong>ms readily recognized as ‘powdery mildews’<br />

because the conidia produced in abundance<br />

on the shoots of infected host plants give<br />

them a whitish powdery appearance. The term<br />

‘powdery mildews’ is often also applied <strong>to</strong> the<br />

organisms causing them. Only angiosperms are<br />

attacked, almost always belonging <strong>to</strong> the dicotyledons.<br />

One notable exception is the powdery<br />

mildew of cereals and grasses caused by Blumeria<br />

graminis (formerly Erysiphe graminis). Other<br />

economically relevant species are Podosphaera<br />

leucotricha causing apple mildew, Podosphaera<br />

(formerly Sphaerotheca) mors-uvae causing<br />

American gooseberry mildew, and Uncinula neca<strong>to</strong>r<br />

(now Erysiphe neca<strong>to</strong>r) causing the powdery<br />

mildew of grapes. Many other species produce<br />

less destructive infections and are ubiqui<strong>to</strong>us in<br />

nature, e.g. Microsphaera alphi<strong>to</strong>ides (now Erysiphe<br />

alphi<strong>to</strong>ides) on the leaves of oak (Quercus spp.) and<br />

Phyllactinia guttata on hazel and many other<br />

broad-leaved trees. Being obligate biotrophs,<br />

powdery mildews cannot be kept in axenic<br />

culture, although Arabi and Jawhar (2002) have<br />

recently grown B. graminis on agar augmented<br />

with shredded barley leaves. The possibility of<br />

infecting Arabidopsis thaliana with several different<br />

powdery mildew species holds promise for<br />

future investigations (Vogel & Somerville, 2002).<br />

In the Erysiphales, the mycelium generally<br />

consists of uninucleate haploid segments. It is<br />

almost always confined <strong>to</strong> the leaf surface, and<br />

infections are limited <strong>to</strong> the epidermal cells<br />

which are penetrated from the outside following<br />

the formation of appressoria. Inside the host cell,<br />

haus<strong>to</strong>ria are formed which provide a large area<br />

of contact with the host. Only relatively few<br />

powdery mildews (e.g. Phyllactinia spp.) are able<br />

<strong>to</strong> penetrate more deeply in<strong>to</strong> the host tissue.<br />

Either way, soon after infection conidia are<br />

produced at the infected leaf surface from a<br />

foot cell, either in basipetal chains or singly as in<br />

Erysiphe if a conidium is released before the next<br />

one is formed. Conidial states of the Erysiphales<br />

are referable <strong>to</strong> the anamorphic genus Oidium,<br />

barring a few specialized genera such as the<br />

Ovulariopsis state of Phyllactinia. A division of<br />

Oidium in<strong>to</strong> subgenera has been proposed (Cook<br />

et al., 1997; Braun et al., 2002; see Fig. 13.1). The<br />

conidia of Erysiphales are generally described as<br />

meristem arthroconidia because the conidiogenous<br />

cell is not homologous <strong>to</strong> a rudimentary<br />

phialide (Hughes, 1953). Numerous crops of<br />

conidia can be produced in a growing season,<br />

and they are the main carriers of infection.<br />

The conidia of Erysiphales are unusual<br />

because unlike most fungal spores they are<br />

fully hydrated. Germination does not require<br />

the uptake of exogenous water and can proceed<br />

even in atmospheres with low relative humidity<br />

(Somers & Horsfall, 1966). In fact, germination


INTRODUCTION<br />

391<br />

Fig13.1 Summary of the key features of the most important genera of Erysiphales.Based on the results of Saenz and Taylor (1999a)<br />

and Braun et al.(2002). Note that Erysiphe is a polyphyletic genus at present.


392 HYMENOASCOMYCETES: ERYSIPHALES<br />

in some species is inhibited by free water. Ultrastructurally,<br />

conidia are highly vacuolated (see<br />

Fig. 13.11). The major carbon and energy reserve<br />

seems <strong>to</strong> be glycogen (McKeen et al., 1967; Roberts<br />

et al., 1996), although lipid droplets have also<br />

been reported, and lipids may contribute about<br />

10% of the dry weight of conidia (Lösel, 1988).<br />

The conidia of Erysiphales are uninucleate.<br />

Ascocarps are usually formed late in the vegetation<br />

period. These have traditionally been<br />

termed cleis<strong>to</strong>thecia, although they differ fundamentally<br />

from those of the Plec<strong>to</strong>mycetes because<br />

the asci of Erysiphales are club-shaped, not globose,<br />

and are formed at one level at the bot<strong>to</strong>m of<br />

the ascocarp rather than being scattered throughout.<br />

Further, at maturity the ascocarp breaks<br />

open by a pre-determined line of weakness,<br />

exposing the asci which forcibly discharge their<br />

spores by a squirt mechanism. In contrast, the<br />

ascospores of Plec<strong>to</strong>mycetes are released passively<br />

when the ascus wall disintegrates. Braun et al.<br />

(2002) have proposed the term chasmothecium<br />

(Gr. chasma ¼ an opening, open mouth) for the<br />

ascocarp of the Erysiphales. Chasmothecia are<br />

brown globose bodies which have no ostiole.<br />

Depending on species, they may contain one or<br />

several asci, and their line of weakness may run<br />

around the equa<strong>to</strong>r of the chasmothecium, or<br />

through its apex. Chasmothecia are often ornamented<br />

by highly characteristic appendages<br />

(Fig. 13.1) which are usually sufficient <strong>to</strong> permit<br />

unambiguous species identification, <strong>to</strong>gether<br />

with the number of asci in the ascocarp (one or<br />

several), the number of ascospores per ascus, and<br />

the identity of the host plant. However, the<br />

phylogenetic value of chasmothecial appendages<br />

appears limited (see Fig. 13.1). In Northern<br />

European climates, the asci are usually fully<br />

formed in late autumn, but chasmothecia do<br />

not open until the following spring when the<br />

host plants begin <strong>to</strong> grow. They are therefore<br />

thought of primarily as overwintering structures,<br />

even if their viability may be low. In countries<br />

with dry hot summers, chasmothecia may serve as<br />

oversummering structures, being formed in late<br />

spring and releasing ascospores in the autumn.<br />

The developmental events taking place during<br />

chasmothecium formation are immensely complex<br />

and have been described by Luttrell (1951)<br />

and Gordon (1966). They are probably similar in<br />

most species. Initially, two superficial uninucleate<br />

hyphae meet and one encircles the other.<br />

The central cell receives a nucleus and enlarges<br />

somewhat. The central cell has been termed the<br />

pseudoascogonial cell because it does not seem<br />

<strong>to</strong> play any direct role in ascus formation. The<br />

pseudoantheridial cells which encircle the pseudoascogonium<br />

divide <strong>to</strong> form the peripheral cells<br />

of the ascocarp. Some outer peripheral cells<br />

(‘mother cells’) develop short septate receptive<br />

hyphae which make contact with vegetative<br />

hyphae on the host surface. Following plasmogamy,<br />

one nucleus is taken up by the receptive<br />

hypha and divides in each segment of the receptive<br />

hypha until one nucleus derived from the<br />

vegetative hypha reaches the mother cell. The<br />

mother cell then divides repeatedly.<br />

At this stage, the immature ascocarp consists<br />

of a pseudoparenchyma<strong>to</strong>us centrum composed<br />

largely of binucleate cells derived from the<br />

mother cells intermixed with some uninucleate<br />

cells, and surrounded by a peridium, some 4 6<br />

cell layers thick. The peridium becomes darkly<br />

pigmented. Uninucleate and binucleate cells<br />

above the middle part of the centrum lyse. Karyogamy<br />

occurs only within certain of the binucleate<br />

cells which are more or less isolated from<br />

the surrounding cells by lysis. These cells then<br />

enlarge <strong>to</strong> form asci. The asci appear <strong>to</strong> grow at<br />

the expense of the uninucleate and binucleate<br />

cells of the centrum, so that eventually the asci<br />

(or a single ascus, depending on the genus)<br />

occupy almost the entire centrum. Meiosis of the<br />

fusion nucleus in developing asci is usually<br />

delayed until the centrum cells have all been<br />

absorbed, although it tends <strong>to</strong> be completed<br />

before the winter dormancy.<br />

13.2 Phylogenetic aspects<br />

The Erysiphales are clearly delimited and defined<br />

as a group, but the question where <strong>to</strong> position<br />

this order within the Ascomycota has aroused<br />

considerable controversy over the past 150 years<br />

or so and is still undecided. Braun et al. (2002)<br />

have given an overview of the taxonomic his<strong>to</strong>ry


BLUMERIA GRAMINIS<br />

393<br />

of the Erysiphales. Affinities with Pyrenomycetes<br />

or Plec<strong>to</strong>mycetes have been proposed in the past<br />

but are not now thought <strong>to</strong> be true. Instead, the<br />

phylogenetic position of the Erysiphales is fairly<br />

isolated, possibly associated weakly with the<br />

Helotiales (Saenz & Taylor, 1999b).<br />

For many years, the taxonomy of genera<br />

and species was based on the system proposed<br />

by Léveillé (1851) who emphasized the features<br />

of the chasmothecium, especially the number of<br />

asci (one or several) and the type of appendage<br />

which may be simple, uncinate (¼ recurved or<br />

hooked), dicho<strong>to</strong>mously branched, or bulbous<br />

(see Fig. 13.1). Recent phylogenetic studies based<br />

on a variety of DNA sequences have revealed that<br />

in the Erysiphales, unlike most other groups of<br />

organisms, the features of asexual reproduction<br />

are more clearly diagnostic than those associated<br />

with sexual reproduction. There is a good correlation<br />

between the anamorphic state and the<br />

groupings obtained by DNA sequence analysis<br />

(Saenz & Taylor, 1999a), and the gross anamorphic<br />

features also correlate with the surface<br />

ornamentations of the conidia as seen by scanning<br />

electron microscopy (Fig. 13.3; Cook et al.,<br />

1997). Thus, the genera of Erysiphales are currently<br />

defined mainly by their anamorphs. In<br />

contrast, the striking chasmothecial appendages<br />

do not correspond well <strong>to</strong> the individual groups<br />

because notably the dicho<strong>to</strong>mous and simple<br />

mycelioid appendages are found in more than<br />

one taxon (Saenz & Taylor, 1999a). One casualty<br />

of these findings is the genus Sphaerotheca with<br />

simple appendages and one ascus per chasmothecium,<br />

which has been incorporated in<strong>to</strong> Podosphaera,<br />

formerly comprising only species with<br />

chasmothecia containing one ascus but bearing<br />

appendages with dicho<strong>to</strong>mously branched tips<br />

(Braun et al., 2002).<br />

We have encountered a great plasticity of<br />

appendages on ascocarps before, especially in the<br />

Plec<strong>to</strong>mycetes where they may be involved in<br />

insect dispersal (see p. 289). In contrast, in the<br />

Erysiphales the chasmothecial appendages seem<br />

<strong>to</strong> be related <strong>to</strong> the type of host infected.<br />

According <strong>to</strong> the analyses of Mori et al. (2000),<br />

the most basal species among the Erysiphales is<br />

Parauncinula septata which has uncinate (hooked)<br />

appendages. This species occurs on oak (Fagaceae),<br />

and Mori et al. (2000) have speculated that the<br />

Erysiphales originated on members of the Fagaceae<br />

because that host family hosts by far the<br />

greatest diversity of powdery mildews (Amano,<br />

1986; Braun, 1987, 1995). Together with another<br />

ancestral species possessing chasmothecia with<br />

uncinate appendages, Caespi<strong>to</strong>theca forestalis,<br />

P. septata may have diverged from other powdery<br />

mildews some 80 90 million years ago<br />

(Takamatsu et al., 2005).<br />

Many of the powdery mildews associated<br />

with the leaves of deciduous trees have hooked<br />

or branched appendages, whereas those on evergreen<br />

or herbaceous plants often have simple<br />

(mycelioid) appendages. It is known that complex<br />

appendages facilitate the attachment of chasmothecia<br />

<strong>to</strong> twigs or the bark of host trees for<br />

overwintering (see p. 407; Gadoury & Pearson,<br />

1988; Cortesi et al., 1995), and it must be advantageous<br />

for the primary inoculum in spring <strong>to</strong> be<br />

close <strong>to</strong> the budding leaves, rather than falling<br />

<strong>to</strong> the ground. Mori et al. (2000) proposed that no<br />

such selection pressure may hold for species parasitizing<br />

herbaceous plants, which might have<br />

permitted a reduction in the complexity of appendages<br />

<strong>to</strong> simple mycelioid ones. Such changes<br />

from woody <strong>to</strong> herbaceous hosts accompanied by<br />

the reduction in the complexity of appendages<br />

may have occurred several times independently<br />

within the Erysiphales (Takamatsu et al., 2000).<br />

13.3 Blumeria graminis<br />

Although Blumeria graminis is somewhat atypical<br />

of a powdery mildew in being a pathogen of<br />

grasses and cereals, we shall describe it in detail<br />

because more is known about it than any other<br />

member of the Erysiphales. Blumeria is separated<br />

from Erysiphe on several grounds. The haus<strong>to</strong>rium<br />

of B. graminis is digitate, i.e. it has striking<br />

finger-like projections (Fig. 13.2c), whereas the<br />

haus<strong>to</strong>ria are knob-like in most members of<br />

the Erysiphales. Upon closer inspection, they<br />

also have lobes, but these are tightly folded<br />

round the main haus<strong>to</strong>rial body, giving it a<br />

globose appearance (Bushnell & Gay, 1978).<br />

A second distinguishing feature of B. graminis


394 HYMENOASCOMYCETES: ERYSIPHALES<br />

Fig13.2 Blumeria graminis. (a) Two-day-old germinating conidium on wheat leaf, showing penetration point surrounded by a‘halo’<br />

(stippled). A haus<strong>to</strong>rium has developed beneath the penetration point. (b) Penetration from an established mycelium. (c) Section of<br />

an epidermal cell showing two penetration points and two haus<strong>to</strong>ria. Note the thickening of the epidermal cell beneath the<br />

penetration point. (d) Mycelium and conidia, showing the swollen flask-shaped foot cell (‘mother cell’).<br />

is that the conidial foot cell is swollen (Fig. 13.2d).<br />

<strong>Third</strong>ly, scanning electron microscopy studies by<br />

Cook et al. (1997) have shown that the surface of<br />

the conidium of B. graminis is distinct from that of<br />

other Erysiphales because of the spiny wall and<br />

the raised septum surrounded by a depressed ring<br />

(Figs. 13.3a,b). These differences are matched by<br />

DNA sequence data (Saenz & Taylor, 1999a).


BLUMERIA GRAMINIS<br />

395<br />

Fig13.3 An example of the use of scanning electron microscopy in fungal taxonomy, separating Blumeria graminis (a,b) from other<br />

members of the Erysiphales (c,d). (a) Whole conidium of B. graminis showing the spiny surface of the wall. (b) The‘annular’ septum<br />

of the conidium.The actual septum is raised and warty, but it is surrounded by a depressed annulus. (c) Conidium of Golovinomyces<br />

(Erysiphe) cynoglossi showing a slightly roughened wall. (d) Conidium of Neoerysiphe (Erysiphe) galeopsidis with its striate conidial wall.<br />

Micrographs kindly provided by R.T. A.Cook. Reprinted from Cook et al. (1997), British Crown copyright1997.<br />

The genus Blumeria contains only one species,<br />

B. graminis, which is separable in<strong>to</strong> several<br />

formae speciales distinguishable by their different<br />

grass (Agropyron, Bromus, Poa) or cereal hosts<br />

(barley, oat, rye, wheat). Thus, B. graminis f. sp.<br />

tritici infects wheat (Triticum) but not barley,<br />

whilst B. graminis f. sp. hordei infects barley<br />

(Hordeum) but not wheat. Blumeria graminis is heterothallic<br />

and hybrids between certain formae<br />

speciales may arise, especially if the hosts themselves<br />

can hybridize. For example, it has been<br />

shown that hybridization between B. graminis ff.<br />

spp. agropyri, tritici and secalis can occur, resulting<br />

in viable ascospores. The three host genera<br />

Agropyron (couch grass), Triticum (wheat) and<br />

Secale (rye) can also hybridize (Hiura, 1978).<br />

Recently, it has become obvious that the definition<br />

of formae speciales is less clear-cut than previously<br />

thought. Extensive overlaps in host spectra<br />

occur especially on wild hosts in the Middle East<br />

where Hordeum spontaneum, the presumed ances<strong>to</strong>r<br />

species of cultivated barley, is endemic and<br />

where B. graminis probably originated (Clarke &<br />

Akhkha, 2002; Wyand & Brown, 2003). Further,<br />

within a given forma specialis, numerous races<br />

can be distinguished by specific features such<br />

as resistance <strong>to</strong> fungicides or ability <strong>to</strong> infect<br />

a given host cultivar.<br />

13.3.1 The infection process<br />

A conidium of B. graminis alighting on the surface<br />

of a grass or cereal leaf will initially come <strong>to</strong> rest<br />

by the tips of its spines. Within seconds of contact<br />

with hydrophobic (but not with hydrophilic)<br />

surfaces, an extracellular matrix is released by<br />

those spines <strong>to</strong>uching the surface. The matrix<br />

contains proteins, including enzymes such as<br />

cutinases and non-specific esterases, and it may<br />

serve in the initial attachment of the conidium <strong>to</strong><br />

the surface (Carver et al., 1999; Nielsen et al., 2000).


396 HYMENOASCOMYCETES: ERYSIPHALES<br />

About 15 min later, a further batch of matrix<br />

material is released from the conidium, but<br />

unlike the initial secretion this second wave<br />

requires de novo protein biosynthesis. Within<br />

2 h, a primary germ tube emerges. In all powdery<br />

mildews except B. graminis, this develops an<br />

appressorium under suitable conditions, but in<br />

B. graminis the primary germ tube grows only <strong>to</strong><br />

a limited distance (up <strong>to</strong> 10 mm). An appressorium<br />

is formed by a separate secondary germ tube<br />

which is much longer than the primary germ<br />

tube (up <strong>to</strong> 40 mm) and becomes septate. If the<br />

primary germ tube fails <strong>to</strong> make contact with<br />

a suitable surface, further short germ tubes may<br />

be emitted. Upon contact with the host surface,<br />

the primary germ tube secretes an adhesive pad<br />

which provides more secure anchorage <strong>to</strong> the<br />

surface. By suspending conidia on a spider’s<br />

thread over different kinds of surface, Carver<br />

and Ingerson (1987) observed that a long appressorial<br />

germ tube is emitted only if the primary<br />

germ tube has made contact with an inductive<br />

surface such as a host leaf. The primary germ tube<br />

therefore functions as a probe. Signals perceived<br />

by the primary germ tube may be the hydrophobicity<br />

of the surface, or minute quantities of<br />

cutin or cellulose degradation products released<br />

by appropriate hydrolytic enzymes which are<br />

secreted by the primary germ tube (Green et al.,<br />

2002). The signals are probably transduced by<br />

cascades containing cAMP and cAMP-dependent<br />

protein kinase A (cPKA), similar <strong>to</strong> those described<br />

in more detail for Magnaporthe grisea (Fig. 12.48).<br />

The primary germ tube may penetrate the surface<br />

of the epidermis <strong>to</strong> a limited extent but does not<br />

achieve successful infection of the host cell. It<br />

may, however, take up water and dissolved substances<br />

from the plant surface (Kunoh & Ishizaki,<br />

1981; Carver & Bushnell, 1983).<br />

Under optimal conditions, the appressoriumforming<br />

germ tube emerges about 3 h after the<br />

primary germ tube. It elongates and its tip<br />

swells <strong>to</strong> form an appressorium which is lobed<br />

(Fig. 13.4b) and non-melanized, in contrast <strong>to</strong><br />

the hemispherical melanized appressorium of<br />

M. grisea described on p. 381. Appressoria<br />

of other species of the Erysiphales may have<br />

different shapes, and these are of taxonomic<br />

significance (see Braun et al., 2002). The sequence<br />

Fig13.4 Penetration events in Blumeria graminis. (a) Conidium<br />

incubated on a cellulose membrane for12 h. After contact of<br />

the primary germ tube (pgt) with the surface, the appressorial<br />

germtube(agt)wasemittedandhasformedanincipient<br />

appressorium. (b) A lobed mature appressorium on the<br />

surface of a wheat leaf. (a) reprinted from Carver et al. (1999),<br />

with permission from Elsevier; original print kindly provided<br />

byT.L.W.Carver.(b)reprintedfromHoward(1997),withkind<br />

permission of Springer Science and Business Media; original<br />

print kindly provided by R. J. Howard.<br />

of infection-related morphogenetic events up <strong>to</strong><br />

this stage is shown in Fig. 13.4.<br />

Penetration is achieved by means of a thin<br />

penetration peg which originates from the underside<br />

of the appressorium. In all probability, both<br />

the activity of secreted wall-degrading enzymes<br />

(cutinases and cellulases) and appressorial turgor<br />

pressure contribute <strong>to</strong> successful penetration<br />

events, with the former predominating (Edwards<br />

& Allen, 1970). Penetration can be achieved within<br />

about 12 h of the conidium landing on the


BLUMERIA GRAMINIS<br />

397<br />

leaf surface. The tip of the penetration peg then<br />

enlarges <strong>to</strong> form the haus<strong>to</strong>rium initial which<br />

differentiates in the course of 2 3 days, invaginating<br />

but not breaching the host plasmalemma.<br />

As a nutrient supply is established, surface<br />

hyphae grow from the appressorium or the<br />

appressorial germ tube and further epidermis<br />

cells are penetrated. It is also possible for the same<br />

host cell <strong>to</strong> be penetrated repeatedly so that it<br />

contains several haus<strong>to</strong>ria. Haus<strong>to</strong>rium formation<br />

represents the end-point of in planta growth;<br />

further penetration occurs from epidermal<br />

hyphae.<br />

13.3.2 Self-defence of the host plant<br />

against infection<br />

Good summaries of this vast <strong>to</strong>pic may be found<br />

in Carver et al. (1995), Giese et al. (1997) and Zeyen<br />

et al. (2002). Although the ungerminated conidium<br />

already emits signals perceived by the epidermis<br />

cell, it is the contact of the primary germ<br />

tube with the host cell surface which elicits initial<br />

defence reactions. These are visible as a dramatic<br />

re-organization of the cy<strong>to</strong>plasm of the attacked<br />

epidermal cell, with dense cy<strong>to</strong>plasm aggregating<br />

beneath the point of contact of the primary germ<br />

tube with the leaf surface. This is followed by<br />

a modification of the epidermal cell wall by<br />

secreted substances. The altered cell wall region is<br />

visible as a halo (Fig. 13.2a). Phenolic substances<br />

and hydrolytic enzymes become incorporated<br />

in<strong>to</strong> the cell wall within the halo region. This is<br />

a non-specific defence reaction because it is elicited<br />

by both virulent and avirulent strains of the<br />

pathogen. A further such halo is produced by the<br />

epidermal cell when the appressorial penetration<br />

peg attempts <strong>to</strong> penetrate. A papilla is then<br />

formed around the penetration peg between the<br />

host cell wall and the host plasma membrane, and<br />

this may succeed in plugging the peg and preventing<br />

infection. Papillae are easily seen with an<br />

epifluorescence microscope because they show<br />

strong au<strong>to</strong>fluorescence, a general indica<strong>to</strong>r of an<br />

attempt by the host plant <strong>to</strong> resist infection.<br />

Au<strong>to</strong>fluorescence is due mainly <strong>to</strong> the accumulation<br />

of phenolic substances which have antimicrobial<br />

activity (von Röpenack et al., 1998).<br />

Additionally, hydrolytic enzymes, callose and<br />

silica may be deposited in the papilla. Papilla<br />

formation is an important mechanism of general<br />

resistance, although it is unknown why some<br />

strains of B. graminis can penetrate the papillae<br />

and others fail. Even if a susceptible host is<br />

infected, only about 70% of the penetration<br />

events succeed beyond the papilla stage. In certain<br />

barley cultivars, particularly thick papillae<br />

are formed because of mutations in which restrictions<br />

of the resistance response are lifted; only<br />

0.5% of infection pegs get through the epidermis<br />

of these mlo mutants, and barley cultivars homozygous<br />

for the mlo allele show broad-spectrum<br />

resistance <strong>to</strong> all strains of B. graminis f. sp. hordei<br />

(Jørgensen, 1994; Collins et al., 2002). It should be<br />

noted here that papillae are not particularly<br />

prominent in infections of dicotyledons by other<br />

powdery mildews.<br />

In cereals attacked by B. graminis, most strainspecific<br />

resistance mechanisms are initiated later,<br />

when the tip of the penetration peg enlarges and<br />

begins <strong>to</strong> differentiate in<strong>to</strong> the first haus<strong>to</strong>rium.<br />

At this point the infected epidermal cell of<br />

a resistant cultivar displays an oxidative burst,<br />

i.e. it releases H 2 O 2 and various enzymes in<strong>to</strong> its<br />

own cy<strong>to</strong>plasm and dies (Zhou et al., 1998). This<br />

phenomenon is known as the hypersensitive<br />

response. Since B. graminis is an obligate biotroph,<br />

penetration ending in a haus<strong>to</strong>rium inside a dead<br />

cell is a wasted effort. The haus<strong>to</strong>rium itself may<br />

also be directly affected by the oxidative burst,<br />

with first signs of degeneration appearing in the<br />

mi<strong>to</strong>chondria (Hippe-Sanwald et al., 1992). If sufficient<br />

nutrient reserves are present in the appressorial<br />

germ tube, further appressoria may be<br />

formed, each resulting in failure <strong>to</strong> establish a<br />

functional haus<strong>to</strong>rium in resistant cultivars.<br />

Successful infection results if the host cell<br />

<strong>to</strong>lerates the establishment of the initial haus<strong>to</strong>rium.<br />

Curiously, therefore, it is sensitivity rather<br />

than <strong>to</strong>lerance which leads <strong>to</strong> resistance. Numerous<br />

genes are involved in the various lines of<br />

defence which barley plants possess against<br />

B. graminis f. sp. hordei (Collinge et al., 2002).<br />

13.3.3 Genetics of plant resistance against<br />

B. graminis<br />

The fact that only certain strains of B. graminis<br />

f. sp. hordei can elicit the hypersensitive response


398 HYMENOASCOMYCETES: ERYSIPHALES<br />

in a given barley cultivar indicates that specific<br />

recognition mechanisms between host and pathogen<br />

must be involved. The molecular basis of<br />

recognition is still obscure, but the genetics are<br />

well unders<strong>to</strong>od. They are based on the genefor-gene<br />

concept first formulated by Flor (1955)<br />

for an interaction between the rust fungus<br />

Melampsora lini and flax, but soon after also<br />

demonstrated for B. graminis and cereal hosts (see<br />

Moseman, 1966). Given that in strain-specific<br />

interactions recognition leads <strong>to</strong> resistance via<br />

the hypersensitive response, every avirulence<br />

gene of the pathogen (e.g. encoding a surface<br />

protein recognized by the host) is matched by<br />

a specific resistance gene of the host (e.g. a<br />

recep<strong>to</strong>r). It follows that avirulence alleles should<br />

be dominant <strong>to</strong> virulence alleles in diploid or<br />

dikaryotic pathogens (not, of course, applicable<br />

<strong>to</strong> the haploid B. graminis), whereas resistance<br />

should be dominant <strong>to</strong> susceptibility in the host.<br />

Successful infection occurs only if the avirulence<br />

gene is modified so that the host can no longer<br />

recognize the pathogen. Numerous resistance<br />

genes have been identified especially in barley<br />

and are being used for breeding programmes,<br />

although it is relatively easy for the pathogen<br />

<strong>to</strong> overcome such single-gene resistance (Brown<br />

et al., 1993; Collins et al., 2002). It should be noted<br />

that there are deviations from the classical genefor-gene<br />

concept in the interaction between<br />

B. graminis f. sp. hordei and barley. Further, interactions<br />

between a given race of B. graminis and its<br />

host can take various courses with intermediates<br />

between complete resistance and full development<br />

of symp<strong>to</strong>ms, due <strong>to</strong> the influence of minor<br />

genes depending on the host’s genetic make-up,<br />

and also due <strong>to</strong> environmental parameters.<br />

A good summary of this complicated <strong>to</strong>pic has<br />

been written by Brown (2002). In general terms,<br />

research on the molecular biology of B. graminis<br />

would benefit greatly from the availability of<br />

a reliable DNA transformation method for this<br />

important pathogen.<br />

13.3.4 The haus<strong>to</strong>rium of B. graminis<br />

In compatible interactions, a functional haus<strong>to</strong>rium<br />

is formed by the enlarging tip of the<br />

penetration peg. The papilla remains as a collar<br />

around the peg at the point where it penetrated<br />

the epidermis wall. The host plasmalemma is<br />

invaginated around the haus<strong>to</strong>rium, but it is not<br />

in direct contact with the plasma membrane of<br />

the haus<strong>to</strong>rium. Instead, the two membranes are<br />

separated by the haus<strong>to</strong>rial wall, and surrounding<br />

it by the extrahaus<strong>to</strong>rial matrix which is of<br />

host origin. It is a compartment with a gelatinous<br />

texture, sealed by the host and pathogen<br />

plasma membranes, and at the epidermal wall by<br />

a collar (Manners, 1989). The host plasmalemma<br />

is strongly modified and seems <strong>to</strong> lack H þ<br />

ATPases, so that the host cell may not be able<br />

<strong>to</strong> control leakage of solutes in<strong>to</strong> the extrahaus<strong>to</strong>rial<br />

matrix (Gay et al., 1987). The escape of<br />

solutes from the extrahaus<strong>to</strong>rial matrix <strong>to</strong> the<br />

cell surface is prevented by the collar seal. Haus<strong>to</strong>ria<br />

of the Erysiphales contain a full complement<br />

of organelles including a single nucleus.<br />

The haus<strong>to</strong>rium is separated from the surface<br />

hypha by a septum which is perforated, thus<br />

permitting nutrient transfer. Ultrastructural<br />

details are summarized in Fig. 13.5 and<br />

have been described by Bracker (1968b) and<br />

Hippe-Sanwald et al. (1992).<br />

Whereas in other biotrophic plant pathogens<br />

nutrients can, <strong>to</strong> a certain extent, be absorbed by<br />

intercellular hyphae in addition <strong>to</strong> haus<strong>to</strong>ria, in<br />

the Erysiphales the haus<strong>to</strong>rium seems <strong>to</strong> be the<br />

sole means of nutrient uptake. Not surprisingly,<br />

the haus<strong>to</strong>rial membrane differs from that of<br />

surface hyphae in terms of protein composition<br />

and physiology (Manners, 1989; Mendgen &<br />

Deising, 1993; Green et al., 2002). Nutrient<br />

uptake has traditionally been studied with the<br />

haus<strong>to</strong>ria of E. pisi which can be isolated intact<br />

from infected pea plants (Gil & Gay, 1977). Sut<strong>to</strong>n<br />

et al. (1999) have shown that glucose is the sugar<br />

which is taken up by haus<strong>to</strong>ria of B. graminis f. sp.<br />

tritici, and that the plant hydrolyses the transport<br />

sugar sucrose before this reaches the epidermal<br />

cells and the haus<strong>to</strong>ria contained within them.<br />

Glucose probably diffuses passively in<strong>to</strong> the<br />

extrahaus<strong>to</strong>rial matrix and is then taken up<br />

across the haus<strong>to</strong>rial membrane by a pro<strong>to</strong>n<br />

uniport mechanism (Sut<strong>to</strong>n et al., 1999). This is<br />

in line with the situation in most other<br />

fungi examined <strong>to</strong> date, which generally take<br />

up glucose but not sucrose (Jennings, 1995).


BLUMERIA GRAMINIS<br />

399<br />

Fig13.5 Interpretation of the<br />

fine structure of the haus<strong>to</strong>rium<br />

of B. graminis f. sp. hordei<br />

(redrawn from Bracker,1968b).<br />

(a) Section of host leaf at point<br />

of penetration.The body of the<br />

haus<strong>to</strong>rium (bo) containing<br />

asinglenucleus(nu)lies<br />

immediately beneath the point<br />

of penetration.The body of the<br />

haus<strong>to</strong>rium is enclosed in a sheath<br />

with extensive matrix (MAT).The<br />

sheath membrane (SM) is close<br />

<strong>to</strong> the host <strong>to</strong>noplast (T).<br />

The sheath membrane bears<br />

invaginations (SI). A single lobe<br />

(lo) of another haus<strong>to</strong>rium<br />

enclosed in an extension of the<br />

sheath is shown <strong>to</strong> the left of the<br />

diagram, and four lobes enclosed<br />

in a common sheath <strong>to</strong> the right.<br />

(b) Enlargement of the neck of a<br />

haus<strong>to</strong>rium. Note the thickened<br />

collar (C) deposited on the host<br />

cell wall (W).The sheath<br />

membrane (SM) is continuous<br />

with the host plasmalemma (E).<br />

Other abbreviations: XW<br />

(cross-wall),CH (channel),<br />

H (host cy<strong>to</strong>plasm).<br />

The difference is that other fungi are capable of<br />

hydrolysing sucrose externally by the activity of<br />

their own secreted invertases.<br />

13.3.5 Life cycle and epidemiology of<br />

B. graminis<br />

About 7 10 days after infection by B. graminis, the<br />

haploid surface mycelium can produce conidia<br />

under field conditions. Conidia develop from<br />

a flask-shaped foot cell within which mi<strong>to</strong>tic<br />

nuclear division occurs. The foot cell elongates<br />

away from the host leaf and a cross-wall cuts off<br />

the hyphal tip. Further cross-walls develop so that<br />

a chain of cells is formed, increasing in length<br />

at its base by further divisions of the foot cell.<br />

Such conidia are said <strong>to</strong> be catenate (Fig. 13.2d).<br />

Each conidium is uninucleate. In successful infections,<br />

a dense stand of conidiophores is produced<br />

so that the lesion appears as a white powdery<br />

pustule (Fig. 13.6a). Aust (1981) has estimated<br />

that a single pustule can release about 1.5 10 4<br />

conidia day 1 . A high spore inoculum can be built<br />

up very quickly, and numerous infection cycles<br />

can occur during one growing season. Dispersal<br />

of powdery mildew conidia is mainly by wind,<br />

whereby short gusts of wind are ideal for spore


400 HYMENOASCOMYCETES: ERYSIPHALES<br />

dispersal (Hammett & Manners, 1971, 1974). This<br />

ties in with the duration of conidium production<br />

(about 3 h for each conidium) and the fact that<br />

usually only the terminal, most mature spore<br />

becomes detached from the conidial chain. Thus,<br />

the observed diurnal rhythm of spore abundance<br />

in the air, with a peak usually in early<br />

afternoon, can probably be explained by fac<strong>to</strong>rs<br />

affecting release, rather than formation, of<br />

spores (Hammett & Manners, 1971, 1974). Both<br />

the formation and release of spores are strongly<br />

inhibited by rain (Hirst, 1953).<br />

The conidia of B. graminis can travel considerable<br />

distances with the prevailing wind. For<br />

instance, Hermansen et al. (1978) have demonstrated<br />

the migration of spores from Northeastern<br />

England and Scotland <strong>to</strong> Denmark, a<br />

journey which would have taken approximately<br />

48 h. In an extensive survey undertaken across the<br />

whole of Europe, Limpert et al. (1999) found that<br />

B. graminis is a nomadic species, with the<br />

prevailing westerly winds driving waves of populations<br />

eastwards at a rate of about 100 km<br />

year 1 . Since such populations encounter hosts<br />

with different spectra of resistance genes, they<br />

face a selective pressure <strong>to</strong> adapt, and this may<br />

explain why the complexity of virulence alleles in<br />

B. graminis populations was found <strong>to</strong> increase<br />

from west <strong>to</strong> east by about one virulence fac<strong>to</strong>r<br />

every 1000 km. Ultra-long distance (intercontinental)<br />

dispersal of viable conidia, as found e.g.<br />

for urediniospores of rust fungi (see p. 632) or<br />

teliospores of smut fungi (p. 639), has not been<br />

reported for powdery mildews.<br />

Whereas conidium formation and release by<br />

B. graminis are favoured by dry windy conditions,<br />

infection proceeds best at 98 100% relative<br />

humidity but is inhibited by free water. The optimum<br />

infection temperature is about 15 20°C.<br />

Thus, the conditions in Northwestern Europe are<br />

ideal for B. graminis in terms of rapidly changing<br />

weather conditions (Smith et al., 1988). Crop<br />

losses of up <strong>to</strong> 40% have been reported, with<br />

barley the most seriously affected cereal. The<br />

timing of infection is important, early infections<br />

reducing the number of ears and later infections<br />

reducing the size of the grain. Crop damage is<br />

due <strong>to</strong> a decrease in pho<strong>to</strong>synthesis in infected<br />

leaves, meaning that they are unable <strong>to</strong><br />

Fig13.6 Blumeria graminis. (a) Conidial pustules on wheat.<br />

(b) Dark spherical chasmothecia nestling in a felt of surface<br />

hyphae on a wheat leaf sheath.<br />

export carbohydrates <strong>to</strong> the developing grains.<br />

Heavily infected leaves may even act as a sink<br />

for pho<strong>to</strong>synthetic product (sucrose) from other<br />

leaves.<br />

The chasmothecia of B. graminis are dark<br />

brown and globose, and nestle in a dense mass<br />

of mycelium formed on the basal leaves and<br />

leaf-sheaths of cereals (Fig. 13.6b). In contrast <strong>to</strong><br />

the chasmothecia of most Erysiphales, those of<br />

B. graminis do not bear any conspicuous thickwalled<br />

appendages (Fig. 13.7a). Each chasmothecium<br />

has a wall made up of several layers of cells,<br />

surrounding a number of asci. At maturity,<br />

it cracks open by swelling of the contents, and<br />

the asci discharge their spores. In Europe,<br />

chasmothecia are formed in late summer and<br />

it is unclear what their exact role in the disease<br />

cycle is. Ascospores of B. graminis formed in the<br />

current season are capable of infection, and<br />

dried chasmothecia can survive under herbarium<br />

conditions for up <strong>to</strong> 13 years (Moseman &<br />

Powers, 1957). However, it is generally assumed


ERYSIPHE<br />

401<br />

that B. graminis survives mainly in the vegetative<br />

state on overwintering host plants, especially<br />

winter varieties of cereals (Jenkyn & Bainbridge,<br />

1978). To survive a single growing season,<br />

B. graminis may have <strong>to</strong> switch four times between<br />

winter barley and summer barley main<br />

crops and their volunteers germinating after<br />

harvest (Limpert et al., 1999). The availability of<br />

living shoots for infection throughout the year is<br />

called the ‘green bridge’. In contrast, in hot and<br />

arid regions such as the Mediterranean, chasmothecia<br />

play a role in the oversummering of the<br />

pathogen on wild grasses such as Hordeum<br />

spontaneum, discharging ascospores when the<br />

host seeds germinate in autumn (Clarke &<br />

Akhkhra, 2002). Since B. graminis seems <strong>to</strong> have<br />

its centre of origin in the Middle East, it is<br />

possible that the chasmothecia are primarily<br />

oversummering rather than overwintering<br />

structures.<br />

13.4 Erysiphe<br />

The genus Erysiphe is currently subject <strong>to</strong><br />

major taxonomic rearrangements. Erysiphe has<br />

traditionally been defined by chasmothecia<br />

with undifferentiated mycelioid appendages<br />

(Fig. 13.8a), and was later distinguished from<br />

Blumeria by its knob-like haus<strong>to</strong>ria and a cylindrical<br />

(non-swollen) conidial foot cell (Fig. 13.8b).<br />

This anamorph gives rise <strong>to</strong> one conidium at a<br />

time, i.e. it is non-catenate. It is called Pseudoidium.<br />

Erysiphe spp. now fall in<strong>to</strong> several branches of<br />

phylogenetic trees, and they are interspersed by<br />

species which have similar anamorphs but chasmothecia<br />

with uncinate or lobed appendages<br />

(Saenz & Taylor, 1999a; Mori et al., 2000). Braun<br />

et al. (2002) therefore proposed the absorption of<br />

genera such as Microsphaera and Uncinula in<strong>to</strong><br />

Erysiphe. The most important species of Erysiphe in<br />

an agricultural context (Spencer, 1978; Smith<br />

et al., 1988) are mentioned below. They seem <strong>to</strong><br />

share the fundamental biological principles of<br />

infecting by wind-dispersed conidia which do not<br />

require or even <strong>to</strong>lerate free water, and causing<br />

disease symp<strong>to</strong>ms as a superficial mycelium<br />

producing haus<strong>to</strong>ria only in epidermal cells.<br />

13.4.1 Erysiphe sensu stric<strong>to</strong><br />

Erysiphe cruciferarum (formerly part of E. polygoni<br />

sensu la<strong>to</strong> or E. communis) is a pathogen of<br />

Fig13.7 Blumeria graminis. (a) Mat of superficial hyphae with<br />

a sectioned chasmothecium containing several asci. (b) Ascus.


402 HYMENOASCOMYCETES: ERYSIPHALES<br />

Fig13.8 Erysiphe polygoni. (a) Chasmothecium showing dark, free-ended equa<strong>to</strong>rial appendages, and the hyaline superficial<br />

mycelium anchoring the chasmothecium <strong>to</strong> the host leaf. (b) T.S. host leaf showing simple haus<strong>to</strong>rium, superficial mycelium and<br />

conidial chain arising from a foot cell which does not appear bulbous.<br />

crucifers, causing severe infections on cabbage<br />

and Brussels sprouts. It is morphologically<br />

similar <strong>to</strong> E. betae which is a specialized pathogen<br />

of sugar beet and related hosts (Beta spp.). Erysiphe<br />

polygoni is now considered a separate species of<br />

no commercial interest, forming powdery<br />

mildew on Rumex and Polygonum. Erysiphe pisi is<br />

a cosmopolitan pathogen of leguminous plants<br />

including peas and lucerne. There are several<br />

formae speciales. This species is of particular<br />

interest because much physiological work has<br />

been done on its haus<strong>to</strong>ria which can be isolated<br />

from infected leaves (Gil & Gay, 1977). Another<br />

powdery mildew fungus on Leguminosae is<br />

Erysiphe trifolii infecting mainly clover and with<br />

formae speciales on Trifolium, Lathyrus, Melilotus<br />

and Lotus. Erysiphe heraclei (¼ E. umbelliferarum)<br />

infects umbellifers such as carrot and celery<br />

and is of limited importance in winter crops in<br />

the Mediterranean. It is exceedingly common<br />

on hogweed (Heracleum sphondylium).<br />

All the above species belong <strong>to</strong> the redescribed<br />

genus Erysiphe, possess a Pseudoidium<br />

anamorph and cluster <strong>to</strong>gether in phylogenetic<br />

analyses (Saenz & Taylor, 1999a; Mori et al., 2000).<br />

13.4.2 Microsphaera and Uncinula<br />

Both these genera group <strong>to</strong>gether with the main<br />

Erysiphe cluster around E. pisi, E. cruciferarum and<br />

other species described above (Saenz & Taylor,<br />

1999a) and will in due course be called Erysiphe<br />

(Braun et al., 2002) possibly with the addition of<br />

further names reflecting distinct phylogenetic<br />

sub-groups rather than characteristics of chasmothecial<br />

appendages. Anamorphic states belong <strong>to</strong><br />

Pseudoidium.<br />

Microsphaera<br />

Chasmothecia of Microsphaera (now Erysiphe sect.<br />

Uncinula) contain several asci, but they carry<br />

appendages with tips showing a highly diagnostic<br />

dicho<strong>to</strong>mous branching (see Fig. 13.10a).<br />

Microsphaera (now Erysiphe) alphi<strong>to</strong>ides is the<br />

cause of oak mildew which is extremely<br />

common, especially on Quercus robur. It appears<br />

from June onwards, infecting mainly leaves<br />

which are produced on sucker shoots and<br />

seedlings, and causing dis<strong>to</strong>rtions <strong>to</strong> growing<br />

shoots until growth finally stalls for the rest of<br />

the growing season (Fig. 13.9). The chasmothecial<br />

stage is not always present but may be more


ERYSIPHE<br />

403<br />

Fig13.9 Summer shoot of oak (Quercus)infectedbyoak<br />

powdery mildew, Microsphaera (now Erysiphe) alphi<strong>to</strong>ides.<br />

Contrast the stunted appearance of the infected foliage with<br />

the healthy basal leaves formed in spring.<br />

common during hot summers or in hot climatic<br />

zones.<br />

Uncinula<br />

Members of the genus Uncinula (now Erysiphe sect.<br />

Uncinula) also produce several asci in each<br />

chasmothecium but are distinguished by their<br />

chasmothecial appendages which are uncinate,<br />

i.e. they have recurved tips. Strikingly similar<br />

appendages are found in the unrelated genus<br />

Sawadaea (see Fig. 13.10b). Undoubtedly the bestknown<br />

species is U. neca<strong>to</strong>r, the cause of powdery<br />

mildew of vines. Chasmothecia are quite rare,<br />

and when the fungus first appeared in Europe<br />

in a glasshouse in Kent in 1845 it was known<br />

only in its anamorphic state and given the name<br />

Oidium tuckeri, after Mr Tucker, the gardener<br />

who discovered it. An accidental introduction<br />

from North America, U. neca<strong>to</strong>r quickly spread<br />

throughout the major vine-growing regions of<br />

Europe, causing such severe damage especially in<br />

France that the entire wine industry was<br />

threatened. By 1854, the French wine production<br />

had fallen from 54 million <strong>to</strong> 10 million hec<strong>to</strong>litres.<br />

Fortunately, Mr Tucker had also discovered<br />

that a mixture of sulphur dust and lime<br />

provided good control of the disease, and this<br />

simple fungicide is still used <strong>to</strong>day (Smith et al.,<br />

1988). An excellent account of the arrival of vine<br />

powdery mildew (and many other fungal<br />

diseases) has been given by Large (1940).<br />

Because of the extremely high value of the<br />

vine crop, extensive epidemiological data have<br />

been compiled and are used for disease forecasting<br />

models (Jarvis et al., 2002). Conidia are,<br />

of course, the major inoculum, and they are<br />

dispersed by wind movements and also by the<br />

air currents generated by high-pressure spraying<br />

equipment (Willocquet & Clerjeau, 1998).<br />

Infections are mainly found on the upper<br />

(adaxial) surface of leaves, in contrast <strong>to</strong> the<br />

grape downy mildew (Plasmopara viticola; see<br />

p. 120) which is more common on the underside<br />

(abaxial surface). The grapes themselves are<br />

also readily infected by U. neca<strong>to</strong>r; if they are<br />

young, they will be aborted al<strong>to</strong>gether whereas<br />

older grapes suffer skin damage, and moulds<br />

such as Botrytis cinerea can easily infect through<br />

these cracks. Uncinula neca<strong>to</strong>r overwinters as<br />

mycelium in dormant buds (Rugner et al., 2002),<br />

although chasmothecia formed in autumn can<br />

survive the winter on twigs and give rise <strong>to</strong><br />

sizeable quantities of ascospores capable of<br />

causing new infections in spring (Jailloux et al.,<br />

1998, 1999).<br />

13.4.3 Species previously attributed <strong>to</strong><br />

Erysiphe<br />

Species of Golovinomyces (formerly Erysiphe cichoracearum<br />

sensu la<strong>to</strong>) are not closely related <strong>to</strong> the<br />

Erysiphe as circumscribed above. They have an<br />

anamorph (Oidium subgenus Reticuloidium; Cook<br />

et al., 1997) that has catenate conidia with slightly<br />

roughened conidial surfaces as seen under the<br />

scanning electron microscope (see Fig. 13.13c).<br />

Braun (1987) has placed the forms affecting<br />

Asteraceae (including lettuce) in G. cichoracearum<br />

and most of the plurivorous forms in the morphologically<br />

very similar G. orontii. Important crops


404 HYMENOASCOMYCETES: ERYSIPHALES<br />

13.5 Podosphaera and Sphaerotheca<br />

These two genera, although easily distinguished<br />

from each other with the light microscope by<br />

their chasmothecial appendages, have now been<br />

grouped <strong>to</strong>gether, and Podosphaera takes precedence<br />

over Sphaerotheca (Braun et al., 2002). Both<br />

produce chasmothecia containing only a single<br />

ascus. The important features of the conidia<br />

of all species grouped here are that they are<br />

catenate and contain conspicuous fibrosin<br />

bodies. These are also seen in some other<br />

genera, e.g. Sawadaea (see p. 405). Fibrosin<br />

bodies are filamen<strong>to</strong>us organelles up <strong>to</strong> 8 mm<br />

long which appear highly light-refractile or<br />

sparkling when viewed with the light microscope<br />

(Fig. 13.11). Their biochemical composition<br />

and function appear <strong>to</strong> be unknown.<br />

Fig13.10 Chasmothecial appendages of Erysiphales.<br />

(a) Flattened dicho<strong>to</strong>mously branched appendages of<br />

Podosphaera clandestina. (b) Branched appendages of Sawadaea<br />

(formerly Uncinula), bicornis with recurved tips.<br />

stated <strong>to</strong> be attacked by the plurivorous species<br />

are those belonging <strong>to</strong> Solanaceae (e.g. <strong>to</strong>bacco)<br />

and Cucurbitaceae (especially melon, cucumber,<br />

squash and pumpkin). However, it is unclear<br />

whether the same species attacks both families.<br />

On cucurbits it has been described as G. cucurbitacearum<br />

but here it seems <strong>to</strong> be of lesser significance<br />

than Sphaerotheca fuliginea (¼ Podosphaera<br />

xanthii) and might have been confused with it in<br />

the past (Jahn et al., 2002). Molecular data indicate<br />

that both Golovinomyces spp. are polyphyletic,<br />

and that the clades do not obviously relate <strong>to</strong><br />

their host ranges. Another catenate species,<br />

Neoerysiphe (formerly Erysiphe) galeopsidis affecting<br />

mainly Lamiaceae differs from Golovinomyces in<br />

having conidia with a minutely striated conidial<br />

surface (Fig. 13.3d).<br />

13.5.1 Sphaerotheca (now Podosphaera)<br />

Common species are S. fuliginea on dandelion<br />

(Taraxacum officinale) and other Asteraceae,<br />

S. macularis causing powdery mildew of hops<br />

(Humulus lupulus), S. mors-uvae (American gooseberry<br />

mildew) and S. pannosa, the common rose<br />

mildew. The chasmothecial structure of Sphaerotheca<br />

closely resembles that of Erysiphe with its<br />

simple appendages, the only notable microscopic<br />

difference being that the chasmothecium contains<br />

only one ascus in Sphaerotheca (Fig. 13.12).<br />

The fine structure of developing and mature<br />

chasmothecia of S. mors-uvae has been studied by<br />

Martin et al. (1976). They have shown that the<br />

darker melanized cells forming the peridium<br />

are, like those of vegetative hyphae, uninucleate.<br />

Most of the inner cells of the chasmothecium are<br />

binucleate, suggesting that they may have arisen<br />

from a binucleate ascogonial fusion cell. Another<br />

interesting discovery was that fibrosin bodies,<br />

previously reported from conidia, are also<br />

present in the ascospores.<br />

The mycelium and conidia of S. pannosa are<br />

common on leaves and shoots of cultivated and<br />

wild roses. Chasmothecia are formed on twigs,<br />

embedded in a dense mycelial felt. Overwintering<br />

is not only by means of ascospores, but<br />

particularly as mycelium within dormant


PHYLLACTINIA AND LEVEILLULA<br />

405<br />

Fig13.11 Sphaerotheca (now Podosphaera) pannosa.(a)Conidial<br />

foot cell producing a chain of conidia. (b) Close-up of mature<br />

conidia. Note the large vacuoles and the elongated<br />

light-refractile fibrosin bodies (arrowheads).<br />

buds, and the chasmothecia may not play any<br />

significant role in perennation. In the case of<br />

S. mors-uvae from blackcurrants, only a small<br />

proportion (


406 HYMENOASCOMYCETES: ERYSIPHALES<br />

Fig13.12 Sphaerotheca (now Podosphaera)<br />

pannosa. (a) Chasmothecium crushed <strong>to</strong> show<br />

a single ascus. (b) Chasmothecium with<br />

discharged ascospores.<br />

gives rise <strong>to</strong> very large and distinctly shaped<br />

conidia.<br />

13.7.1 Phyllactinia<br />

Phyllactinia guttata (¼ P. corylea) grows on the<br />

leaves of hazel (Corylus avellana) and other woody<br />

plants, forming chasmothecia on the lower leaf<br />

surface in late summer and autumn. In this<br />

species the superficial mycelium produces short<br />

lateral appressoria which emit penetration<br />

hyphae through the s<strong>to</strong>mata in<strong>to</strong> the mesophyll<br />

(Fig. 13.13b). Directly beneath the s<strong>to</strong>ma, the<br />

penetration peg swells <strong>to</strong> produce a subs<strong>to</strong>matal<br />

vesicle from which septate hyphae grow in<strong>to</strong> the<br />

leaf mesophyll, penetrating mesophyll cells and<br />

forming haus<strong>to</strong>ria within them. The conidia are<br />

club-shaped and are formed singly (Fig. 13.13a).<br />

However, in a damp chamber pseudo-chains of<br />

up <strong>to</strong> four conidia may hang <strong>to</strong>gether. Conidia<br />

of this type are classified in the anamorphic<br />

genus Ovulariopsis.<br />

Phyllactinia guttata presents a most intriguing<br />

spore dispersal mechanism which is easily demonstrated<br />

in mycology classes because the same<br />

hazel trees are reliably infected each year and,<br />

once a source has been located, leaves can be


PHYLLACTINIA AND LEVEILLULA<br />

407<br />

Fig13.13 Phyllactinia guttata.<br />

(a) Conidiophores showing the<br />

single terminal conidium. (b) T.S.<br />

leaf of Corylus avellana showing<br />

penetration of s<strong>to</strong>ma in lower<br />

epidermis, formation of a<br />

subs<strong>to</strong>matal vesicle (arrow) and<br />

extension of the mycelium in<strong>to</strong><br />

the mesophyll.<br />

collected and kept dry until required for classes<br />

(Weber & Webster, 2001a). Chasmothecia are<br />

formed only on the lower leaf surface. They bear<br />

two types of appendage, an equa<strong>to</strong>rial group of<br />

radiating unbranched appendages with bulbous<br />

bases, and a crown of highly branched appendages<br />

which secrete mucilage (Fig. 13.14). The base<br />

of the bulbous appendage is thin-walled in the<br />

region facing the leaf surface, but thick-walled<br />

facing away from the leaf. On drying, the<br />

appendage bends <strong>to</strong>wards the leaf surface as<br />

the thin part buckles inwards, and the pressure<br />

of the appendage tip levers the chasmothecium<br />

free from the superficial mycelium (for a video<br />

sequence, see Webster, 2006b). The bulbous<br />

appendages now function as ‘flights’ or vanes,<br />

and the chasmothecium plummets downwards<br />

like a shuttlecock, orientated during its fall with<br />

the sticky mucilage on the lower side. The blob of<br />

mucilage between the apical crown of branched<br />

appendages helps <strong>to</strong> stick the chasmothecium<br />

on<strong>to</strong> twigs and leaves. The asci usually contain<br />

only two spores (Fig. 13.14e). Overwintered ascocarps<br />

open by means of an equa<strong>to</strong>rial line of<br />

dehiscence, and the base of the ascocarp hinges<br />

back <strong>to</strong> place the asci in a suitable position for<br />

ascus discharge <strong>to</strong> occur (Cullum & Webster,<br />

1977; Weber & Webster, 2001a).<br />

For his<strong>to</strong>rical interest, we may note that in<br />

1861 the Tulasne brothers, working on P. guttata<br />

and other powdery mildews, were among the first<br />

<strong>to</strong> demonstrate and illustrate in timeless beauty<br />

the connection between a conidial state and a<br />

morphologically very different-looking sexual<br />

state (see Tulasne & Tulasne, 1931).<br />

13.7.2 Leveillula<br />

This genus is related <strong>to</strong> Phyllactinia. Its members<br />

are common in warmer countries and occasionally<br />

in glasshouses in cooler areas. Leveillula<br />

taurica attacks a wide range of mainly herbaceous<br />

plants including <strong>to</strong>ma<strong>to</strong> and pepper. It is


408 HYMENOASCOMYCETES: ERYSIPHALES<br />

distinguished from Phyllactinia by its rare chasmothecia<br />

bearing only mycelioid appendages<br />

and by its spear-shaped conidia (Oidiopsis<br />

anamorph).<br />

13.8 Controlofpowderymildew<br />

diseases<br />

Several different strategies are employed <strong>to</strong> control<br />

powdery mildew diseases in various crops,<br />

chemical control and breeding for resistance<br />

being by far the most important in commercial<br />

terms. We describe them here in detail because<br />

the fundamental principles apply <strong>to</strong> diseases<br />

caused by many other groups of fungi within<br />

the Eumycota, as mentioned elsewhere in this<br />

book.<br />

13.8.1 Breeding for resistance<br />

Wheat and barley were among the first crop<br />

plants for which resistance breeding programmes<br />

were initiated almost a century ago,<br />

following Biffen’s (1905) seminal work on the<br />

genetics of resistance <strong>to</strong> wheat stripe rust,<br />

Puccinia striiformis. Initial attempts at resistance<br />

breeding against B. graminis as well as other<br />

biotrophic plant pathogens used major gene<br />

resistance based on introducing single resistance<br />

genes in<strong>to</strong> the cultivar of choice. Numerous such<br />

major resistance genes originating from cereal<br />

cultivars as well as wild grasses related <strong>to</strong> wheat<br />

or barley are now available for breeding purposes<br />

(Hsam & Zeller, 2002). Breeding for major gene<br />

resistance has also been pursued in many other<br />

crops susceptible <strong>to</strong> powdery mildew diseases,<br />

such as melons and cucumbers (Podosphaera<br />

xanthii and Golovinomyces cichoracearum; Jahn<br />

et al., 2002), or clover (Erysiphe trifolii), hops<br />

(Sphaerotheca macularis) and gooseberries<br />

(Sphaerotheca mors-uvae) (see Smith et al., 1988).<br />

Breeding for resistance in slow-growing perennial<br />

crops such as apple or vines is obviously<br />

rather more difficult.<br />

Major gene resistance is usually based on<br />

a recognition mechanism involving the hypersensitive<br />

response (see pp. 397 398). The danger<br />

inherent in major gene resistance breeding is<br />

that the pathogen can overcome resistance by<br />

mutation of its corresponding avirulence allele<br />

<strong>to</strong> virulence. This is a frequent occurrence, e.g. in<br />

B. graminis. On the other hand, the frequency of<br />

virulence alleles in the field may decrease again<br />

in pathogen populations no longer exposed <strong>to</strong><br />

the cultivar and its resistance gene. Further, by<br />

co-ordinating the release of resistant cultivars,<br />

the effect of major gene resistance in crop<br />

protection can be maximized. For instance, the<br />

‘green bridge’ of B. graminis (see p. 399) can be<br />

broken if the summer and winter cereal cultivars<br />

sown in any one year carry different resistance<br />

genes. Further, it is possible <strong>to</strong> combine several<br />

resistance genes in one host variety, a process<br />

known as ‘pyramiding’. This requires the pathogen<br />

<strong>to</strong> develop multiple virulence alleles before<br />

it can infect such a crop variety (Hsam & Zeller,<br />

2002). The danger, of course, lies in the creation<br />

of multiply resistant ‘super-races’ of B. graminis<br />

or other powdery mildews.<br />

An interesting type of resistance in barley is<br />

mediated by the mlo allele which, as mentioned<br />

on p. 397, mediates the formation of very thick<br />

papillae, restricting infection by all races of<br />

B. graminis f. sp. hordei. This resistance is unusual<br />

for a broad-spectrum resistance in giving<br />

almost <strong>to</strong>tal control, and it is exceptional for<br />

a single-gene resistance in that it has remained<br />

stable in barley cultivars since it was<br />

introduced about 20 years ago (Jørgensen, 1994;<br />

Collins et al., 2002). In 1990, 30% of the spring<br />

barley acreage was sown with mlo resistant<br />

barley cultivars (Jørgensen, 1992). More<br />

commonly, broad-spectrum resistance is based<br />

on the combined effects of several minor genes<br />

and it is only partial, i.e. it retards infection<br />

and sporulation of the pathogen but does not<br />

al<strong>to</strong>gether prevent disease. It is sometimes<br />

termed horizontal resistance because it controls<br />

many different races of the pathogen <strong>to</strong> the<br />

same limited level, in contrast <strong>to</strong> vertical<br />

resistance mediated by major resistance genes<br />

in which individual races are controlled <strong>to</strong>tally,<br />

and others not at all. Although the molecular<br />

basis of most resistance genes is still unknown,<br />

it is noteworthy that some genes involved in<br />

horizontal resistance become more effective in


CONTROL OF POWDERY MILDEW DISEASES<br />

409<br />

Fig13.14 Phyllactinia guttata. (a) Chasmothecium on lower side of hazel (Corylus) leaf.The radiating bulbous appendages are<br />

horizontal.The branched secre<strong>to</strong>ry appendages which crown the chasmothecium are on the morphologically upper side, i.e. the<br />

side <strong>to</strong> which the ascus apices point. (b) Position of chasmothecium during its fall from the host leaf.The bulbous appendages are now<br />

folded <strong>to</strong> form flight vanes, ensuring that the sticky mass of mucilage faces downwards. (c) Diagrammatic representation of an open<br />

chasmothecium.The chasmothecium is shown attachedby mucilage <strong>to</strong> a surface.The chasmothecium has openedby a circumscissile<br />

line of weakness, and has hinged back so that the apices of the asci now point outwards. Arrows indicate the direction of ascospore<br />

discharge. (d) A bulbous appendage, showing the differential thickening of the wall of the bulb.Collapse of the thinner walls result<br />

in movement of the appendage. (e) Two-spored ascus. (f) Branched secre<strong>to</strong>ry appendage. (a,b) <strong>to</strong> same scale; (d f) <strong>to</strong> same scale.


410 HYMENOASCOMYCETES: ERYSIPHALES<br />

adult plants. This phenomenon is called adult<br />

plant resistance (Hsam & Zeller, 2002). Thus,<br />

it can be envisaged that the products of horizontal<br />

resistance genes may be involved, for<br />

example, in producing a thicker cuticle, cell<br />

wall or papilla.<br />

Resistance breeding will certainly remain one<br />

of the key methods for controlling powdery<br />

mildews, and it is likely that major and minor<br />

genes will be combined <strong>to</strong> produce cultivars<br />

with more durable and effective resistance <strong>to</strong><br />

B. graminis and other powdery mildews. Further,<br />

genetic engineering may enable breeders <strong>to</strong> produce<br />

cultivars with overexpresssed pathogenesisrelated<br />

(PR) genes (Salmeron et al., 2002). The<br />

principles of PR genes are explained briefly on<br />

p. 116.<br />

13.8.2 Chemical control<br />

A good overview of fungicides against powdery<br />

mildews has been given by Holloman and<br />

Wheeler (2002). By far the oldest remedy against<br />

powdery mildew is powdered elemental sulphur,<br />

which was mentioned by Homer in about<br />

1000 BC (Agrios, 2005) and was rediscovered in<br />

the nineteenth century and combined with lime<br />

for enhanced efficacy (Large, 1940). Sulphur<br />

lime mixtures saved the French wine industry<br />

from ruin in the mid nineteenth century when<br />

U. neca<strong>to</strong>r appeared in Europe, just as Bordeaux<br />

mixture (based on copper sulphate and lime)<br />

played a crucial role in protecting vines against<br />

Plasmopara viticola later that century (see p. 120).<br />

In the early twentieth century, as described<br />

above, efforts <strong>to</strong> control B. graminis on<br />

cereals shifted <strong>to</strong> breeding for resistance.<br />

Dithiocarbamate (see Fig. 5.27) was released in<br />

1934 and, like sulphur, had a purely protectant<br />

activity.<br />

The first systemic fungicide against powdery<br />

mildews was the benzimidazole benomyl<br />

(Fig. 13.15a) which is converted <strong>to</strong> the active<br />

molecule carbendazim inside the plant. Carbendazim<br />

binds <strong>to</strong> tubulin proteins, interfering<br />

with their assembly in<strong>to</strong> microtubules and<br />

especially with the nuclear spindle during<br />

mi<strong>to</strong>sis (Davidse & Ishii, 1995). Resistance against<br />

benomyl is common among plant-pathogenic<br />

fungi and is usually due <strong>to</strong> a mutation in the<br />

b-tubulin gene at the site where carbendazim<br />

binds, thereby reducing its affinity (Davidse &<br />

Ishii, 1995).<br />

The next two important groups of systemic<br />

fungicides <strong>to</strong> be released against powdery mildews<br />

were the morpholines (Pommer, 1995) and<br />

2-aminopyrimidines introduced in the 1960s<br />

(Hollomon & Schmidt, 1995). Both are used more<br />

or less exclusively against powdery mildews.<br />

Fig13.15 <strong>Fungi</strong>cides against powdery mildews. (a) The benzimidazole benomyl.The active substance, carbendazim, is<br />

produced in planta by removal of the side chain at the position indicated by the arrow. (b) The morpholine tridemorph. (c) The<br />

2-aminopyrimidine ethirimol. (d) The triazole triadimefon.The molecule becomes reduced <strong>to</strong> its more active alcohol in planta and<br />

by fungi at the position indicated by the arrow. (e) Strobilurin A, an antifungal substance from Strobilurus tenacellus. (f) Kresoxim<br />

methyl, the first strobilurin-based fungicide. (g) Quinoxyfen. (h) Benzothiadiazole (¼ benzo(1,2,3)thiadiazole-7-carbothioic acid<br />

S-methyl ester), an inducer of systemic acquired resistance.


CONTROL OF POWDERY MILDEW DISEASES<br />

411<br />

Fig13.16 One of several possible sterol biosynthesis pathways<br />

in fungi. Most fungi have ergosterol as their major membrane<br />

sterol, but in the Erysiphales it is ergosta-5,24(28)-dien-3b-ol<br />

(Loeffler et al.,1992). Putative targets of various sterol<br />

biosynthesis inhibi<strong>to</strong>rs are indicated as follows. (1) Allylamines,<br />

e.g. terbinafine (against Candida albicans;seeFig.10.9).<br />

(2) Triazoles (against fungi pathogenic <strong>to</strong> humans as well as<br />

plants). A specific demethylation step is inhibited at C14.<br />

(3) Morpholines, e.g. fenpropimorph (against plant-pathogenic<br />

fungi).These compounds seem <strong>to</strong> inhibit two different steps<br />

in sterol biosynthesis. Re-drawn and modified from<br />

Kerkenaar (1995).<br />

Morpholines (e.g. tridemorph; Fig. 13.15b) were<br />

the first sterol biosynthesis inhibi<strong>to</strong>rs, acting by<br />

inhibiting two sterol-modifying enzymes in fungi<br />

(Fig. 13.16) whilst not affecting sterol biosynthesis<br />

in plants (Kerkenaar, 1995; Uesugi, 1998). A more<br />

recent fungicide, fenpropimorph, is derived<br />

from tridemorph and has a similar mode of<br />

action but is used against other fungi as well.<br />

In contrast, the 2-aminopyrimidines (Fig. 13.15c)<br />

act by inhibiting the incorporation of adenine<br />

in<strong>to</strong> nucleic acids, thereby halting DNA synthesis.<br />

The effects are visible early, usually at the stage of<br />

formation of the first haus<strong>to</strong>rium (Holloman &<br />

Schmidt, 1995).<br />

Triazoles, which we have already encountered<br />

as drugs in the treatment of Candida<br />

infections (Fig. 10.9), are also extensively used<br />

in agriculture because of their selective action in<br />

sterol biosynthesis. There is a huge diversity of<br />

structures (Kuck et al., 1995; Uesugi, 1998), but<br />

they all seem <strong>to</strong> act as inhibi<strong>to</strong>rs of a cy<strong>to</strong>chrome<br />

P-450 involved in demethylating sterols at the C-<br />

14 position (Fig. 13.16). These compounds are<br />

therefore collectively called demethylation inhibi<strong>to</strong>rs<br />

(DMIs). The target organism is unable <strong>to</strong><br />

produce a functional membrane. Triazoles are<br />

active against a very wide range of fungal<br />

pathogens and are systemic fungicides with<br />

curative properties. One of the first important<br />

examples against powdery mildews and other<br />

biotrophic plant pathogens was triadimefon,<br />

which is reduced in plants and fungi <strong>to</strong> its<br />

more active alcohol (Fig. 13.15d). However,<br />

resistance has arisen on numerous occasions,<br />

and various resistance mechanisms have been


412 HYMENOASCOMYCETES: ERYSIPHALES<br />

implicated, such as exclusion or reduced uptake<br />

of the fungicides, altered membrane lipid<br />

composition with reduced sterol content, or<br />

mutation of the fungicide-binding site on the<br />

cy<strong>to</strong>chrome P-450 enzyme.<br />

One of the most important recently introduced<br />

classes of fungicides are the strobilurins.<br />

They are based on strobilurin A (Fig. 13.15e),<br />

a natural product from the basidiomycete<br />

Strobilurus tenacellus initially described by Anke<br />

et al. (1977) as a strongly and selectively antifungal<br />

substance. This was derivatized <strong>to</strong> give<br />

kresoxim methyl, the first commercial strobilurin-type<br />

fungicide (Fig. 13.15f). The mode of<br />

action is based on an inhibition of complex III<br />

of the mi<strong>to</strong>chondrial respira<strong>to</strong>ry chain (Anke,<br />

1997). Resistance of Strobilurus tenacellus <strong>to</strong> its<br />

own product is based on a mutation of three<br />

amino acids in the target polypeptide, and resistance<br />

based on similar mechanisms has arisen<br />

in many fungal plant pathogens. In consequence,<br />

strobilurins are not currently recommended for<br />

control of powdery mildews on cereals and<br />

cucumber/melon crops (Hollomon & Wheeler,<br />

2002).<br />

A recently released compound with exclusive<br />

activity against powdery mildews is quinoxyfen<br />

(Fig. 13.15g). The mode of action is interesting<br />

because this compound selectively inhibits<br />

morphogenetic events related <strong>to</strong> infection, such<br />

as germination or appressorium formation, but<br />

not vegetative growth. It is therefore non-<strong>to</strong>xic <strong>to</strong><br />

other fungi. Quinoxyfen is not systemic but is<br />

distributed in the vapour phase and binds <strong>to</strong><br />

the surface waxes of the epidermis. It is thus<br />

ideally placed for activity against the germinating<br />

powdery mildew conidium (Hollomon &<br />

Wheeler, 2002). Non-fungicidal compounds<br />

with such a highly specific mode of action<br />

have fewer side effects against non-target organisms<br />

such as saprotrophic or mycorrhizal fungi,<br />

and are likely <strong>to</strong> increase in importance in<br />

future.<br />

Activa<strong>to</strong>rs of systemic acquired resistance<br />

(SAR; see p. 116), especially the salicylic acid<br />

derivative benzothiadiazole (Fig. 13.15h), may<br />

prove their worth in the prevention of powdery<br />

mildew infections as well as many other plant<br />

diseases (Salmeron et al., 2002), although they<br />

cannot cure existing infections. Since SAR activa<strong>to</strong>rs<br />

trigger the expression of a multitude of<br />

defence-related proteins and other mechanisms<br />

in the crop plant, powdery mildews are unlikely<br />

<strong>to</strong> develop resistance against them, like they<br />

have done against most of the currently used<br />

fungicides, many of which target a single site in<br />

the physiology of the pathogen. Therefore, fungicides<br />

must be used as cocktails containing<br />

chemicals with different modes of action, and<br />

following strict recommendations (Hollomon &<br />

Wheeler, 2002).<br />

13.8.3 Biological control<br />

Their exposed habitat on the leaf surface renders<br />

powdery mildews potentially susceptible <strong>to</strong> parasitism<br />

by phylloplane fungi. Perhaps the bestknown<br />

of them is Ampelomyces quisqualis, which<br />

produces conidia within pycnidial fruit bodies<br />

associated with the hyphae, conidia and, less<br />

frequently, chasmothecia of powdery mildews.<br />

It is commonly found in nature (Falk et al., 1995;<br />

Kiss, 1997). Attempts are being made <strong>to</strong> develop<br />

it as a biological control agent (see Bélanger &<br />

Labbé, 2002), but success is limited by the ability<br />

of powdery mildews <strong>to</strong> grow at lower humidities<br />

than A. quisqualis. This may be alleviated by<br />

the application of the fungus in paraffin oil<br />

(Bélanger & Labbé, 2002). One positive aspect is<br />

that A. quisqualis is <strong>to</strong>lerant of several fungicides<br />

used against powdery mildews, so that integrated<br />

control is possible in principle (Sundheim,<br />

1982).<br />

In general, however, the biological control of<br />

powdery mildews, like that of most other<br />

airborne pathogens, would appear <strong>to</strong> be limited<br />

<strong>to</strong> greenhouses because there the environment<br />

can be controlled <strong>to</strong> a certain extent. With powdery<br />

mildews, humidity appears <strong>to</strong> be the crucial<br />

parameter not only in interactions with<br />

A. quisqualis, but also other potential control<br />

fungi such as Verticillium lecanii or basidiomycete<br />

yeasts, e.g. Tilletiopsis and Pseudozyma (summarized<br />

in Bélanger & Labbé, 2002). All these, in<br />

contrast <strong>to</strong> A. quisqualis, control powdery<br />

mildews through the production of biologically<br />

active substances rather than by direct parasitism.<br />

Fatty acid derivatives seem <strong>to</strong> be the main<br />

active compounds, and these may act by


Plate1 Slimemoulds(Myxomycota).(a)White coralloidsporocarps ofCeratiomyxafruticulosa (Pro<strong>to</strong>steliomycetes) onrotting wood.<br />

(b) Developing aethalia of Lycogala epidendron on rotting wood. (c) Aethalium of Reticularialycoperdon with a silvery grey peridium.<br />

(d) Aethalium of R.lycoperdonwithitsperidiumruptured <strong>to</strong>reveala darkbrownpowderymass of spores. (e) Arcyriadenudata onrotting<br />

wood.The sporangia have openedup, releasing their dullred spores.The capillitium network is exposed. (f) Phaneroplasmodium of<br />

Physarumpolycephalum producing numerous stalked sporangia on an agar surface. (g) Slightly immature aethalium of Fuligo septica.<br />

(h) Clustered stalked sporangia of Stemonitisaxifera. (e) and (h) kindly providedby G.L.Barron.


Plate 2 Oomycota. (a) Salmon infected by Saprolegnia sp. (b) Damping-off of cress caused by Pythium sp. (c) Sudden death of a<br />

20 -year-old Pseudotsuga menziesii plantation caused by Pythium undulatum (¼Phy<strong>to</strong>phthora undulata) due <strong>to</strong> root rot following heavy<br />

summer rains. (d) Phy<strong>to</strong>phthora rootrotsymp<strong>to</strong>msofAbies procera. (e) Late blight symp<strong>to</strong>ms on pota<strong>to</strong> leaf caused by Phy<strong>to</strong>phthora<br />

infestans. (f) Pink rot of pota<strong>to</strong> tuber caused by Phy<strong>to</strong>phthora erythroseptica. Killed tissue shows a pink discoloration after cutting and<br />

exposure <strong>to</strong> air for 30 min. (g) Peronospora parasitica on wallflower (Cheiranthus cheiri) showing the dis<strong>to</strong>rted host shoot and<br />

downy appearance of sporulating regions. (h) Cauliflower leaf infected with Albugo candida, showing the typical white blister rust<br />

symp<strong>to</strong>ms caused by the eruption of sporangia through the epidermis.


Plate 3 Chytridiomycota (a c) and Zygomycota (d i). (a,b) Synchytrium endobioticum. (a) Excavated pota<strong>to</strong> plant showing a gall at<br />

the base of a shoot. (b) Leafy gall stage at the soil surface. (c) Synchytrium taraxaci, sporangial sori on dandelion (Taraxacum officinale).<br />

(d) Strawberry infected by Rhizopus s<strong>to</strong>lonifer. (e) Sporangiophores of Spinellus fusiger on an old basidiocarp of Mycena pura.<br />

(f) Sporangiophore of Piloboluscrystallinus producing a hemispherical black sporangium. Ayellow band below the base of the swollen<br />

subsporangial vesicle, the ocellus, is enriched in carotenoids.The stalk and vesicle bear droplets of liquid. (g) En<strong>to</strong>mophthora muscae.<br />

Dead fly attached <strong>to</strong> a window pane, surrounded by a halo of discharged conidia. (h) Furia americana. Dead blowflies attached <strong>to</strong> a<br />

leaf.Conidiophores have penetrated between the abdominal segments. (i) Sporocarp of the pea truffle Endogone lactiflua,cu<strong>to</strong>pen<br />

<strong>to</strong> reveal the zygospores.


Plate 4 Archiascomycetes (a c) and Plec<strong>to</strong>mycetes (d f). (a) Peach leaf infected withTaphrina deformans.(b)Taphrina betulina<br />

causing witches’ broom disease on birch. (c) Taphrina amen<strong>to</strong>rum causing abnormal enlargement of individual catkin segments on<br />

Alnus glutinosa.(d)Aspergillus parasiticus. Petri dish with parasexual recombinants deficient in melanin biosynthesis leading <strong>to</strong><br />

colourless spores, and/or in afla<strong>to</strong>xin biosynthesis which leads <strong>to</strong> the accumulation of bright yellow or orange pigments. (e) Orange<br />

rotted by Penicillium digitatum. (f) Excavated cleis<strong>to</strong>thecium of Elaphomyces infected by Cordyceps ophioglossoides.Germinating<br />

ascospores have accumulated as whitish pustules around the perithecial ostioles. (d) kindly provided by J.F. Peberdy, (f) by J. Benn.


Plate 5 Pyrenomycetes. (a) Daldinia concentrica on an old log of ash.One stroma has been cut open. (b) Stromata of Xylaria longipes<br />

on a buried sycamore trunk. (c) Hypocrea pulvinata, perithecial crust on the underside of an old fruit body of Pip<strong>to</strong>porus betulinus.<br />

(d) Nectria cinnabarina with conidial pustules and dark red perithecial stromata. (e) Orange-coloured sporodochia of Fusarium<br />

heterosporum parasitizing sclerotia of Claviceps purpurea var. spartinae on Spartina anglica. (f) Perithecial stroma of Epichloe typhina<br />

‘choking’ a shoot of Dactylisglomerata.(g)Cordyceps militaris, perithecial stromata emerging from a buried insect pupa. (h) Dutch elm<br />

disease. (i) Anthracnose on broad bean caused by Colle<strong>to</strong>trichum lindemuthianum. (c) kindly provided by J. Benn.


Plate 6 Apothecia of operculate discomycetes (Pezizales). (a) Pyronema domesticum fruiting on au<strong>to</strong>claved pottery. (b) Aleuria<br />

aurantia.(c)Peziza vesiculosa on freshly manured garden soil. (d) Ascobolus furfuraceus on agar.The purple dots represent ripe asci.<br />

(e) Sarcoscypha australis on an old log of ash (Fraxinus). (f) Helvella crispa.(g)Morchella esculenta. (g) kindly provided by P. Davoli.


Plate 7 Inoperculate Discomycetes (Helotiales). (a,b) Sclerotinia curreyana. (a) Sclerotia formed inside culms of Juncus effusus.<br />

(b) Stalked apothecia of S. curreyana arising from a sclerotium in spring. (c) Rutstroemia echinophila on the shell of edible chestnuts.<br />

(d) Mollisiacinerea, a common saprotroph on twigs. (e) Dasyscyphusvirgineus producing small apothecia with a hairy margin on rotting<br />

wood. (f) Leotia lubrica, a saprotroph soil fungus with stalked apothecia. (g) Chlorociboria aeruginascens on a decaying sycamore twig.<br />

Sections of the colonized wood show its green discolouration. (h) Bisporellacitrina on a beech twig. (i) Bulgariainquinans forming black<br />

gelatinous apothecia on freshly felled oak trunks. (j) Apothecial stromata of Cyttaria darwinii on living twigs of the southern beech<br />

Nothofagus pumilio. (d) and (e) kindly provided by H. Anke, (f) by P. Davoli. (a,b) reprinted from Weber and Webster (2003), with<br />

permission from Elsevier.


Plate 8 Thalli of lichens. (a) The crus<strong>to</strong>se lichen Lecanora muralis. Numerous apothecia are seen in the centre of the thallus. (b) The<br />

crus<strong>to</strong>se Rhizocarpon geographicum. Its mosaic-like appearance is due <strong>to</strong> greenish-yellow vegetative thalli with small dark apothecial<br />

areas, and larger black prothalli. (c) The crus<strong>to</strong>se foliose lichen Xanthoria parietina.The yellow colour is due <strong>to</strong> the pigment parietin.<br />

Numerous apothecia are seen in the centre of the thallus. (d) Peltigeracanina on mossy boulders.The foliose fleshy thallus is coloured<br />

dark blue green by the cyanobacterium Nos<strong>to</strong>c. Note the white rhizinae and the reddish apothecial areas. (e) Cladonia floerkiana.<br />

Fruticose upright podetia arise from a squamulose horizontal thallus. (f) The reindeer lichen Cladina rangiferina growing among<br />

ground vegetation. (g) Usnea florida, a fruticose lichen hanging down from tree branches. (b) kindly provided by B. Bu«del.


Plate 9 Fruit bodies of Homobasidiomycetes: euagarics and bole<strong>to</strong>id clades. (a) Amanita muscaria, probably the best-known of all<br />

mushrooms. (b) Amanita caesarea, a prized delicacy in Mediterranean countries. (c) The scarlet waxcap, Hygrocybe coccinea. (d) The<br />

winter fungus, Flammulina velutipes.(e)Pholiota squarrosa.(f)Boletus erythropus.(g)Suillus granulatus.(h)Boletus badius. Healthy fruit<br />

body (left) and others attacked by the mycoparasitic mould, Sepedonium chrysospermum. (e) kindly provided by P. Davoli.


Plate10 Fruitbodies of Homobasidiomycetes: other clades. (a) Trametes versicolor, the artist’s fungus or turkey tail on a birchbranch.<br />

Note the bleached (white-rot) appearance of the wood at the broken end. (b) Laetiporussulphureus, the chicken of the woods.<br />

(c) Lactariusdeliciosus.(d)Chondrostereum purpureum, cause of silver-leaf disease on plum trees. (e) Thelephoraterrestris,an<br />

ec<strong>to</strong>mycorrhizalspecies.(f)Phellinusigniarius fruitingon an oldwillow tree. (g) Cantharelluscibarius,thechanterelle.(h)Ramariabotrytis.


Plate11 Gasteromycetes (a f) and Heterobasidiomycetes (g i). (a) Calvatia excipuliformis.(b)Pisolithus tinc<strong>to</strong>rius, one gasterocarp<br />

cut open <strong>to</strong> reveal the peridioles. (c) Rhizopogon sp.; excavated gasterocarp which has been cut open. (d) Phallus impudicus,the<br />

common stinkhorn. (e) Clathrusruber.(f)Aseroe rubra.(g)Calocera viscosa.(h)Auricularia auricula-judae.(i)Tremella mesenterica.<br />

(f) reprinted from Fuhrer (2005), with permission by Bloomings Books Pty Ltd. Image kindly provided by B.Fuhrer.


Plate12 Urediniomycetes (a g) and Ustilaginomycetes (h j). (a) Aeciospore of Puccinia distincta.Lipid droplets have been displayed<br />

by the two nuclei in the upper spore. (b) Aecial infection of Pucciniacaricina on stinging nettle. (c) Uredinia (orange pustules) and telia<br />

(dark purple lesions) of Phragmidiumviolaceum on theunderside of a leaf ofbramble. (d f)Gymnosporangiumfuscum. (d) Telial horns on<br />

aswollencankeronJuniperus in spring. (e) Spermogonial infection on a pear leaf in midsummer. (f) Roestelioid aecia on the underside<br />

of a pear leaf in autumn.The aecial caps are connected <strong>to</strong> the aecialbase by trellis-like threads. (g) Uredinia of Melampsora sp. on<br />

Populus tremula. (h) Maize smutcausedby Ustilago maydis. Swollen kernels have become convertedin<strong>to</strong> teliospore-bearing tumours.<br />

(i) Exobasidium sp. on an ornamental Azalea.Infectedleaves are strongly hypertrophied. (j) Systemic Exobasidiumvaccinii infection of<br />

blueberry (Vaccinium myrtillus).The infected shoot (left) shows a reddish discolouration. (d) and (i) kindly providedby H.Weber.


CONTROL OF POWDERY MILDEW DISEASES<br />

413<br />

inserting themselves in<strong>to</strong> the powdery mildew<br />

plasma membrane, disrupting its structural<br />

integrity (Avis & Bélanger, 2002; Urquhart &<br />

Punja, 2002). Other explanations for the strongly<br />

inhibi<strong>to</strong>ry effect of specific fatty acids (especially<br />

cis-monounsaturated ones) against powdery mildews<br />

are, however, also possible (Wang et al.,<br />

2002). Unsaturated fatty acids could also be the<br />

basis for the anti-powdery mildew activity of<br />

cow’s milk described by Bettiol (1999).


14<br />

Hymenoascomycetes: Pezizales (operculate<br />

discomycetes)<br />

14.1 <strong>Introduction</strong><br />

The order Pezizales contains the operculate<br />

discomycetes which are the most readily recognized<br />

cup fungi. The order is large, containing<br />

some 15 families, about 160 genera and over<br />

1100 species (Kirk et al., 2001). Most are terrestrial<br />

and saprotrophic on soil, burnt ground, decaying<br />

wood, compost or dung, but some form sheathing<br />

mycorrhiza (ec<strong>to</strong>mycorrhiza) with trees<br />

(Maia et al., 1996). A somewhat exceptional<br />

case is Rhizina undulata, which causes root rot<br />

of conifers in plantation situations, usually starting<br />

from areas affected by recent fires (Callan,<br />

1993). Whilst most species of Pezizales produce<br />

epigeous fruit bodies above ground level and<br />

have active ascus discharge mechanisms with<br />

wind-dispersed ascospores, the truffles (e.g. Tuber<br />

and Terfezia) form subterranean (hypogeous)<br />

ascomata. The dispersal of truffles relies on<br />

the ripe ascomata being eaten by rodents and<br />

other mammals attracted by their strong odour.<br />

The ascospores survive digestion and defaecation.<br />

There are also aquatic Pezizales, growing on<br />

wood in streams or other wet places. An overview<br />

of the Pezizales may be found in Pfister and<br />

Kimbrough (2001). Keys <strong>to</strong> genera are given<br />

by Korf (1972) and Dissing et al. (2000).<br />

The ascocarp is generally an apothecium<br />

(p. 245) which can range in diameter from less<br />

than one millimetre <strong>to</strong> several centimetres. It is<br />

often cup-shaped or disc-like, fleshy, sometimes<br />

stalked, and frequently brightly coloured.<br />

The asci are, in most cases, cylindrical with<br />

a well-defined lid called operculum (see<br />

pp. 239 241) which is the characteristic feature<br />

of the Pezizales. Members of this order are<br />

therefore often referred <strong>to</strong> as ‘operculate discomycetes’.<br />

The asci are interspersed by filamen<strong>to</strong>us<br />

paraphyses, the tips of which often contain<br />

carotenoids giving the apothecia their striking<br />

yellow, orange or red colours. Several unusual<br />

carotenoids are known in nature only from<br />

apothecia of Pezizales (Gill & Steglich, 1987).<br />

The ascus wall appears distinctly two-layered<br />

under the light microscope, but the two layers do<br />

not separate during ascospore discharge as they<br />

do in functionally bitunicate (i.e. fissitunicate)<br />

ascomycetes (see p. 239). The ascus of the<br />

Pezizales is thus bitunicate but non-fissitunicate.<br />

In many Pezizales the ascus wall is amyloid, i.e. it<br />

stains blue or purple with Melzer’s iodine<br />

(an aqueous solution of iodine and KI). The<br />

blue-staining properties are associated with an<br />

outer mucilaginous layer which may extend for<br />

the whole length of the ascus or may be confined<br />

<strong>to</strong> an apical region (Samuelson, 1978a,b). The<br />

presence or absence of the amyloid staining<br />

property may help in distinguishing certain<br />

genera (see Hansen et al., 2001). The septal pore<br />

formed at the base of the ascus has characteristic<br />

features which may be useful in classification<br />

(Kimbrough, 1994). The ascospores are colourless<br />

<strong>to</strong> reddish-brown, globose <strong>to</strong> ellipsoidal, unicellular<br />

(i.e. non-septate) and may be uninucleate,


PYRONEMA (PYRONEMATACEAE)<br />

415<br />

Table 14.1. Families of the Pezizales which are commonly encountered in nature. Data from Kirk et al. (2001).<br />

Families printed in bold are considered in more detail in this chapter.<br />

Family Number of species Examples<br />

Ascobolaceae (p.419) 118 Ascobolus, Saccobolus<br />

Discinaceae 25 Gyromitra<br />

Helvellaceae (p.423) 68 Helvella<br />

Morchellaceae (p.427) 38 Morchella,Verpa<br />

Pezizaceae (p.419) 160 Peziza<br />

Pyronemataceae (p.415) 462 Aleuria,Otidea, Pyronema, Scutellinia<br />

Rhizinaceae 2 Rhizina<br />

Sarcoscyphaceae 36 Sarcoscypha<br />

Sarcosomataceae 31 Galiella, Urnula<br />

Terfeziaceae 15 Terfezia<br />

Tuberaceae (p.423) 87 Tuber<br />

quadrinucleate or multinucleate. They contain<br />

one or several large lipid globules and may have<br />

smooth or ornamented walls. Individual asci<br />

may discharge their spores asynchronously, or<br />

large numbers of asci may shoot off their spores<br />

simultaneously. In this case, the spores may be<br />

released in a visible cloud in a process known as<br />

‘puffing’ (see Fig. 8.14). Buller (1934) investigated<br />

this phenomenon in detail and reported that<br />

ascospore puffing produces an audible hissing<br />

sound. The asci of truffles, in contrast, are saclike<br />

or globose, with no functional operculum,<br />

and the spores are released passively.<br />

The fruit bodies of some members of the<br />

group are highly prized culinary delicacies, notably<br />

truffles (Tuber spp.) and morels (Morchella<br />

spp.). Gyromitra esculenta, the false morel, was<br />

formerly also widely consumed but, following<br />

a series of often fatal mushroom poisonings,<br />

was eventually found <strong>to</strong> contain the heat-labile<br />

<strong>to</strong>xin gyromitrin. This is readily converted<br />

in<strong>to</strong> hydrazine derivatives such as the rocket<br />

fuel methylhydrazine, which are highly <strong>to</strong>xic<br />

and carcinogenic (Bresinsky & Besl, 1990). Illustrations<br />

of the ascomata of commonly occurring<br />

Pezizales have been provided by Breitenbach<br />

and Kränzlin (1984) and Dennis (1981). Selected<br />

examples are presented on Plate 6.<br />

Molecular phylogenetic analyses indicate<br />

that the Pezizales are probably a primitive<br />

group ancestral <strong>to</strong> other Euascomycete orders<br />

(Fig. 8.17; Gargas & Taylor, 1995; Landvik<br />

et al., 1997). This implies that apothecia are<br />

an ancient type of ascoma, with cleis<strong>to</strong>thecia<br />

and perithecia representing later developments.<br />

Whilst the Pezizales as a whole are monophyletic,<br />

the arrangement in<strong>to</strong> families within this order<br />

is still tentative (Harring<strong>to</strong>n et al., 1999). The more<br />

important of the currently recognized families<br />

are listed in Table 14.1. Since the morphological<br />

and ecological features of pezizalean fungi<br />

often transcend the family boundaries, we will<br />

describe a few characteristic genera in more<br />

detail by their biological features, merely<br />

indicating their family assignment where<br />

appropriate.<br />

14.2 Pyronema (Pyronemataceae)<br />

The apothecia of Pyronema develop on burnt<br />

soil and on heat-sterilized composts in glasshouses.<br />

There are two species, P. omphalodes (¼<br />

P. confluens) and P. domesticum (Moore & Korf,<br />

1963). In P. omphalodes the apothecia are confluent<br />

and lack marginal hairs, whilst in<br />

P. domesticum the apothecia are more discrete,<br />

and surrounded by tapering hairs (Fig. 14.1a).<br />

Pyronema domesticum forms sclerotia in culture<br />

(Moore, 1962), whilst P. omphalodes does not.<br />

In earlier studies the distinction between the<br />

two species was sometimes not appreciated and


416 HYMENOASCOMYCETES: PEZIZALES (OPERCULATE DISCOMYCETES)<br />

Fig14.1 Pyronema domesticum.<br />

(a) Apothecium showing hymenium<br />

and excipular hairs. (b) Group of<br />

ascogonia and antheridia. (c) V.S.<br />

through developing apothecium<br />

showing several ascogonia<br />

producing ascogenous hyphae, and<br />

the development of paraphyses and<br />

excipulum from the ascogonial<br />

stalks. (d) Enlarged view of the<br />

ascogonium and developing<br />

ascogenous hyphae. (e j) Stages<br />

in the development of asci.<br />

(e) Binucleate tip of ascogenous<br />

hypha beginning <strong>to</strong> form a crozier.<br />

(f) Quadrinucleate stage.<br />

(g) Septation of crozier <strong>to</strong> form<br />

a binucleate penultimate cell.<br />

(h) Development of ascus from<br />

binucleate cell. (i) Completion of<br />

first meiotic division.<br />

(j) Completion of second meiotic<br />

division. Note the proliferation of<br />

a new ascogenous hypha from the<br />

stalk cell.<br />

some reports purporting <strong>to</strong> be on P. confluens<br />

may well have been based on P. domesticum.<br />

These earlier studies include classical accounts<br />

of the cy<strong>to</strong>logy of the development of ascogenous<br />

hyphae, croziers and asci (see Moore, 1963).<br />

14.2.1 Development of asci in Pyronema<br />

Both species of Pyronema are homothallic and<br />

grow rapidly in agar culture or on sterilized soil<br />

and within 4 5 days form pink apothecia about<br />

1 2 mm in diameter (Plate 6a; Webster & Weber,<br />

2001). Apothecia of P. domesticum arise from<br />

clusters of ascogonia and antheridia formed<br />

by repeated dicho<strong>to</strong>my of a single hypha. The<br />

ascogonia are fatter than the antheridia and<br />

each ascogonium is surmounted by a tubular<br />

recurved trichogyne which grows <strong>to</strong> make contact<br />

with the tip of an antheridium (Fig. 14.1b).<br />

Both antheridia and ascogonia are multinucleate<br />

and, following fusion of the trichogyne with the<br />

antheridium by breakdown of the walls separating<br />

them (plasmogamy), antheridial nuclei<br />

stream in<strong>to</strong> the ascogonium and each antheridial<br />

nucleus becomes paired with an ascogonial<br />

nucleus. Nuclear fusion (karyogamy) does not<br />

occur at this stage, but the paired nuclei remain<br />

associated with each other. Branched investing<br />

hyphae develop from the ascogonial stalks and<br />

envelop the cluster of fertilized ascogonia, ultimately<br />

making up the tissues of the apothecium,<br />

i.e. the medullary and ectal excipulum. Several<br />

ascogenous hyphae extend from each ascogonium<br />

and grow between the surrounding<br />

investing hyphae (Figs. 14.1c,d).<br />

Further development follows the common<br />

ascomycete pattern outlined in Fig. 8.10 (I. M.<br />

Wilson, 1952; Hung & Wells, 1971). The ascogenous<br />

hyphae are branched and septate at their<br />

tips, which recurve <strong>to</strong> form croziers. The tip of<br />

the crozier is binucleate and the two nuclei


ALEURIA (PYRONEMATACEAE)<br />

417<br />

simultaneously divide by mi<strong>to</strong>sis (conjugate<br />

mi<strong>to</strong>sis). Two septa cut off a uninucleate terminal<br />

cell, a binucleate penultimate cell and<br />

a uninucleate antepenultimate cell (the stalk<br />

cell). The binucleate penultimate cell is the ascus<br />

mother cell and the two nuclei within it fuse,<br />

i.e. karyogamy now occurs. The diploid fusion<br />

nucleus undergoes meiosis and the four resulting<br />

haploid nuclei then divide mi<strong>to</strong>tically so<br />

that eight haploid nuclei result around which<br />

the eight ascospores are subsequently cleaved<br />

(Reeves, 1967). No further nuclear divisions occur<br />

so that each ascospore contains one haploid<br />

nucleus. No special inclusions are seen in the<br />

septa of the crozier, but electron-dense plugs are<br />

formed at the base of the ascus (Hung & Wells,<br />

1971; Kimbrough, 1994). When the uninucleate<br />

terminal cell grows backwards and makes<br />

contact with the stalk cell, their walls break<br />

down and a new binucleate cell is formed which<br />

grows on <strong>to</strong> form a further crozier and another<br />

ascus, a process which is repeated so that a single<br />

ascogenous hypha may produce several asci<br />

(Fig. 14.2b). The ascus mother cell elongates and<br />

acquires a cylindrical shape. It is surrounded by<br />

filamen<strong>to</strong>us paraphyses. These develop from the<br />

stalks of the ascogonia (I. M. Wilson, 1952) and<br />

also appear <strong>to</strong> arise from ascogenous hyphae<br />

(Fig. 14.2c). As the ascus matures it extends above<br />

the layer of paraphyses and explodes, throwing<br />

out its ascospores. The operculum may persist<br />

as a hinged lid (Fig. 14.2c) or may be blown<br />

off. During the development of asci, before<br />

the ascospores are cleaved out, the operculum<br />

becomes apparent as a thickened rim of wall<br />

material at the upper end of the ascus.<br />

Mature ascospores have three wall layers,<br />

a thicker, electron-transparent inner layer (the<br />

endospore), a thinner electron-opaque epispore,<br />

and an outer fibrous perispore of variable thickness.<br />

The perispore lies immediately within<br />

the ascospore-investing membrane, and, as ascospores<br />

mature, this membrane continues <strong>to</strong><br />

produce vesicles, leading <strong>to</strong> degradation of the<br />

perispore. At discharge, the ascospores do not<br />

stick <strong>to</strong>gether but remain separate from each<br />

other (Hung, 1977). Merkus (1976) has described<br />

the development and structure of the ascospores<br />

of P. omphalodes in similar terms.<br />

Much is known about the conditions under<br />

which P. domesticum forms apothecia and sclerotia<br />

(Moore-Landecker, 1975, 1992). Light is<br />

required for apothecium development, with<br />

white, blue and far-red light being particularly<br />

effective. Sclerotium formation is inhibited by<br />

intense blue light.<br />

14.2.2 Ecology of Pyronema<br />

In nature, the apothecia of both species of<br />

Pyronema are among the first <strong>to</strong> appear on<br />

burnt ground following volcanic eruptions, wildfires,<br />

controlled burns and bonfires. Pyronema<br />

forms part of a characteristic group of ‘phoenicoid<br />

fungi’, i.e. fungi arising from ashes. Many<br />

other operculate discomycetes are also phoenicoid<br />

(Carpenter & Trappe, 1985; Dix & Webster,<br />

1995). The ascospores of P. domesticum germinate<br />

readily at 20°C, although a short exposure <strong>to</strong><br />

50°C enhances germination. Apothecium formation<br />

is inhibited by the presence of other<br />

soil-inhabiting organisms and it is possible that<br />

the preference for burnt ground and steamsterilized<br />

soil is associated with its rapid growth<br />

and inability <strong>to</strong> compete with other soil biota<br />

(El-Abyad & Webster, 1968a,b).<br />

Unnatural (i.e. man-made) situations in which<br />

P. domesticum fruits are steam-sterilized soils<br />

and composts used in horticulture, and plaster<br />

prepared by slaking lime, a process which generates<br />

heat. Supposedly sterile surgical gauzes<br />

manufactured from Chinese cot<strong>to</strong>n have been<br />

found <strong>to</strong> be contaminated with P. domesticum<br />

due <strong>to</strong> insufficient radiation treatment during<br />

manufacture. The source is likely <strong>to</strong> be raw<br />

cot<strong>to</strong>n materials possibly already contaminated<br />

in the field (Yan, 1998). Labora<strong>to</strong>ry experiments<br />

have shown that the g-irradiation resistance<br />

of P. domesticum ascospores is higher even than<br />

that of Bacillus endospores (Richter & Barnard,<br />

2002).<br />

14.3 Aleuria (Pyronemataceae)<br />

Aleuria seems <strong>to</strong> be closely related <strong>to</strong> Pyronema<br />

(Landvik et al., 1997). There are about 10 species<br />

of Aleuria, growing especially on forest soil.


418 HYMENOASCOMYCETES: PEZIZALES (OPERCULATE DISCOMYCETES)<br />

Fig14.2 Pyronemadomesticum. (a) An immature ascus (left).The ascogenous hypha fromwhich itdeveloped continues <strong>to</strong> proliferate.<br />

(b) More magnified view of the tip of an ascogenous hypha showing repeated proliferation.The three stippled cells represent<br />

penultimate cells of croziers probably destined <strong>to</strong> develop in<strong>to</strong> asci. (c) Mature asci, one discharged and showing an operculum. A<br />

paraphysis is also shown apparently arising from the ascogenous hypha.


ASCOBOLUS (ASCOBOLACEAE)<br />

419<br />

Aleuria is similar in appearance <strong>to</strong> Peziza (see<br />

below) but is distinguished from it in having<br />

non-amyloid asci. The best-known species is A.<br />

aurantia, the so-called orange-peel fungus which<br />

forms strikingly orange-coloured cup-shaped<br />

apothecia from about 1 10 cm in diameter in<br />

woodland and grassland soil during autumn<br />

(Plate 6b). Iso<strong>to</strong>pic analyses of ascocarps using<br />

15 N and 13 C iso<strong>to</strong>pes indicate that the fungus<br />

may be mycorrhizal (Hobbie et al., 2001). The<br />

orange colour of the hymenium is due <strong>to</strong><br />

carotenoid-enriched granules in the club-shaped<br />

tips of the paraphyses (Fig. 14.3). The main<br />

carotenoids are b-carotene, g-carotene and aleuriaxanthin<br />

(Gill & Steglich, 1987).<br />

The ascospores contain two lipid globules and<br />

the ascospore wall is ornamented by a honeycomb-like<br />

series of raised ridges which represent<br />

secondary wall material. It is derived from<br />

granules in the perisporic sac which condense<br />

in<strong>to</strong> larger spherical dense bodies. These accumulate<br />

and become attached <strong>to</strong> the epispore<br />

(Merkus, 1976; Wu & Kimbrough, 1993).<br />

14.4 Peziza (Pezizaceae)<br />

Peziza is a large genus containing around<br />

100 species. Analyses based on a combination<br />

of morphological and molecular evidence indicate<br />

that this genus is polyphyletic, i.e. it<br />

contains a number of unrelated taxa which<br />

have been grouped <strong>to</strong>gether artificially. Peziza<br />

in its traditional sense can be differentiated<br />

in<strong>to</strong> at least eight clades (Hansen et al., 2005).<br />

Hohmeyer (1986) has provided a key <strong>to</strong> European<br />

species. For descriptions and illustrations see<br />

Dennis (1981), Breitenbach and Kränzlin (1984)<br />

and Dissing et al. (2000). The apothecia are cupshaped,<br />

often large (2 5 cm or more), usually<br />

pale brown and fleshy (Plate 6c). They are<br />

commonly encountered in a very wide range<br />

of habitats including soil, manure heaps, dung,<br />

rotting wood or straw, burnt ground and sand<br />

dunes. About six species have hypogeous ascocarps<br />

(Trappe, 1979). The ascus wall is, in general,<br />

amyloid. In P. succosa the blue-staining by I 2 /KI<br />

is confined <strong>to</strong> the ascus apex (Samuelson, 1978a).<br />

Fig14.3 Aleuria aurantia. Asci, ascospores and paraphyses.<br />

The tips of the paraphyses are filled with granules containing<br />

orange-coloured carotenoids.<br />

The conidial states of Peziza spp., where known,<br />

have been classified in the anamorph genus<br />

Oedocephalum (Fig. 14.4).<br />

A genus with more strikingly coloured apothecia<br />

is Sarcoscypha (Plate 6e). Although similar<br />

in appearance <strong>to</strong> Peziza, these two genera are not<br />

closely related (Landvik et al., 1997).<br />

14.5 Ascobolus (Ascobolaceae)<br />

There are about 80 species of Ascobolus<br />

(van Brummelen, 1967). Most of them are


420 HYMENOASCOMYCETES: PEZIZALES (OPERCULATE DISCOMYCETES)<br />

Fig14.4 Oedocephalum conidial state of Peziza subviolacea<br />

(¼ P. praetervisa). (a) Conidiophores terminating in a<br />

club-shaped vesicle bearing numerous dry blas<strong>to</strong>conidia.<br />

(b) Details of developing conidia (<strong>to</strong>p) and a vesicle from which<br />

the conidia have been detached (bot<strong>to</strong>m). (c) Conidia, two of<br />

which are germinating. Scale bar ¼ 40mm (a,b) and 20 mm(c).<br />

From Webster et al. (1964), with permission from Elsevier.<br />

coprophilous, growing on the dung of herbivorous<br />

animals, but A. carbonarius grows on old<br />

bonfire sites. Common coprophilous species<br />

are A. furfuraceus (¼ A. stercorarius) which is very<br />

commonly found on old cattle dung, often<br />

along with A. immersus (Figs. 14.5 and 14.6,<br />

respectively). Whilst these species are heterothallic,<br />

some others, e.g. A. crenulatus (¼ A. viridulus),<br />

are homothallic. Characteristic features of all<br />

species are the purple colour of the ascospores<br />

and the protruding, operculate asci. Ascobolus<br />

furfuraceus forms yellowish saucer-shaped apothecia<br />

up <strong>to</strong> 5 mm in diameter, and when mature<br />

the surface of the apothecium is studded with<br />

purple dots which mark the ripe asci (Plate 6d).<br />

As the asci mature they elongate above the<br />

general level of the hymenium. The ascus tips<br />

are pho<strong>to</strong>tropic and this ensures that when they<br />

explode the spores are thrown away from<br />

the dung. The ascospores have a mucilaginous<br />

perispore which aids attachment. Ascobolus<br />

immersus has yellow globose apothecia about<br />

1 2 mm in diameter, with very large ascospores<br />

(about 70 30 mm). The perispores cause all<br />

the eight ascospores <strong>to</strong> adhere <strong>to</strong> form a single<br />

projectile about 250 mm long, capable of being<br />

discharged for up <strong>to</strong> 30 cm horizontally. In<br />

general, multi-spored projectiles have a lower<br />

surface-<strong>to</strong>-volume ratio and are projected further<br />

than single spores (see p. 317 for Sordaria).<br />

There is a general trend among coprophilous<br />

fungi <strong>to</strong>wards multi-spored projectiles. In the<br />

genus Saccobolus, which also belongs <strong>to</strong> the<br />

Ascobolaceae, all eight spores are firmly cemented<br />

<strong>to</strong>gether by their perispores.<br />

The spores of Ascobolus become attached <strong>to</strong><br />

herbage and, when eaten by a herbivore, germinate<br />

in the faeces. It is likely that digestion<br />

stimulates spore germination. Most spores fail<br />

<strong>to</strong> germinate on nutrient media but can be<br />

triggered <strong>to</strong> do so by treatment with 0.4% NaOH<br />

or bile salts, and incubation at 37°C. The purple<br />

pigment in the spore wall develops late and is<br />

deposited within the perispore from the ascus<br />

epiplasm. Immature spores are colourless. The<br />

spore wall bears longitudinal colourless striations<br />

in some species, e.g. A. immersus (Fig. 14.5)<br />

and A. furfuraceus (Fig. 14.6). Both species can be<br />

grown and induced <strong>to</strong> form apothecia in culture<br />

(see Webster & Weber, 2001).<br />

14.5.1 Mating behaviour<br />

There is variation in the mating behaviour of<br />

different species of Ascobolus. A single ascospore<br />

culture of A. scatigenus (¼ A. magnificus) does not<br />

produce apothecia. Sex organs (coiled ascogonia<br />

and antheridia) are formed only when mycelia<br />

of different mating types are grown <strong>to</strong>gether.<br />

Each strain is hermaphroditic, i.e. is capable<br />

of developing both ascogonia and antheridia.<br />

However, A. scatigenus is self-incompatible,<br />

i.e. the antheridia of one strain do not fertilize<br />

the ascogonia borne on the same mycelium.<br />

The ascospores of this fungus are of two types, A


ASCOBOLUS (ASCOBOLACEAE)<br />

421<br />

Fig14.5 Ascobolus immersus.(a)Apothecium<br />

showing two projecting asci. Immature asci can<br />

be seen below the general level of the surface.<br />

A single projectile consisting of eight adhering<br />

ascospores is shown above the apothecium.<br />

Note the operculum which has also been<br />

projected. (b) Tip of ripe ascus showing the<br />

operculum. (c) Tip of discharged ascus. In this<br />

case the operculum has remained attached <strong>to</strong><br />

the ascus tip.<br />

and a, and fertilization can only occur between<br />

an A ascogonium and an a antheridium, or<br />

vice versa. There is thus a gene for mating<br />

type represented in two idiomorphs A and a,<br />

and incompatibility is controlled by this gene<br />

irrespective of the presence of both types of<br />

sex organ on each strain. There is no morphological<br />

difference between the two different<br />

mating type strains.<br />

A similar situation occurs in A. furfuraceus,<br />

but here there are no antheridia. Instead, each<br />

strain at first produces chains of arthrospores<br />

or oidia (see Fig. 14.6c). The oidia can germinate<br />

<strong>to</strong> form a fresh mycelium, i.e. they can function<br />

asexually as conidia, but they also play a part<br />

in sexual reproduction. Mites and flies may<br />

transport oidia of one strain <strong>to</strong> the mycelium<br />

of the opposite strain, and following this,<br />

apothecia develop. The process of fertilization<br />

has been studied by Bistis (1956, 1957) and Bistis<br />

and Raper (1963). If an A oidium is transferred<br />

<strong>to</strong> an a mycelium, the oidium fails <strong>to</strong> germinate<br />

and within 10 h an ascogonial primordium<br />

appears on the a mycelium (Fig. 14.6d). The<br />

ascogonium consists of a broad coiled base and<br />

a narrow apical trichogyne which shows chemotropic<br />

growth <strong>to</strong>wards the oidium and eventually<br />

fuses with it. There is evidence that this sequence<br />

of events is under hormonal control, and it<br />

has been suggested that a fresh A oidium is not<br />

immediately capable of inducing development<br />

of ascogonial primordia, but must itself at first<br />

be sexually activated by a messenger secreted<br />

by the a mycelium. Following activation, the<br />

oidium can induce ascogonial development.<br />

By substitution experiments, it has been shown<br />

that an A ascogonium can be induced <strong>to</strong> fuse<br />

with an A oidium, i.e. an oidium of the same<br />

mating type, but apothecia fail <strong>to</strong> develop<br />

from such fusions. In compatible crosses, fertile<br />

apothecia develop within about 10 days of fertilization,<br />

each ascus producing four A and four<br />

a spores.<br />

In A. immersus there are no morphologically<br />

distinguishable antheridia and there are<br />

no oidia. When A and a mating type mycelia<br />

are grown <strong>to</strong>gether in culture, multinucleate<br />

ascogonia develop which are fertilized by fusion<br />

with slender multinucleate hyphae of the<br />

opposite strain. Experimentally, fertilization<br />

can also be achieved using homogenized mycelial<br />

fragments of one strain <strong>to</strong> spermatize a<br />

strain of opposite mating type (Lewis & Decaris,<br />

1974).


422 HYMENOASCOMYCETES: PEZIZALES (OPERCULATE DISCOMYCETES)<br />

Fig14.6 Ascobolus furfuraceus. (a) Group of asci and<br />

paraphyses.One ascus is mature and contains<br />

purple-pigmented ascospores. (b) The same ascus as<br />

shown in a after discharge.The ascus has decreased in size<br />

during discharge. Note the operculum. (c) Arthrospores<br />

(oidia) developed in five-day-old culture. (d) Coiled<br />

ascogonium formed in a single ascospore culture within<br />

48 h of adding oidia of the opposite mating type.<br />

The trichogyne of the ascogonium has grown <strong>to</strong>wards<br />

the oidium and has fused with it.<br />

14.5.2 Apothecium development<br />

The development of apothecia of A. furfuraceus<br />

has been studied by various authors, including<br />

Wells (1970, 1972) and O’Donnell et al. (1974).<br />

The fertilized ascogonium becomes surrounded<br />

by sheath hyphae which develop from the<br />

ascogonial stalk, and the paraphyses and excipular<br />

tissues develop from the sheath hyphae.<br />

The ascogonium gives rise <strong>to</strong> numerous ascogenous<br />

hyphae. Van Brummelen (1967) has<br />

distinguished two kinds of ascocarp development<br />

in Ascobolus. In gymnohymenial forms,<br />

the hymenium is exposed from the beginning<br />

until the maturation of the asci. In cleis<strong>to</strong>hymenial<br />

forms the hymenium is enclosed during<br />

its early development. Ascobolus furfuraceus and<br />

A. immersus are examples of cleis<strong>to</strong>hymenial<br />

development.<br />

14.5.3 Ascosporogenesis<br />

The details of ascosporogenesis have been<br />

studied in A. immersus by Wu and Kimbrough<br />

(1992) and in A. stic<strong>to</strong>ideus by Wu and Kimbrough<br />

(2001). Although A. immersus has smooth ascospores<br />

and A. stic<strong>to</strong>ideus has ornamented spores,<br />

development is very similar. As in most ascomycetes,<br />

invagination of the ascus vesicle encloses<br />

the young ascospores by a two-layered sporedelimiting<br />

membrane. The inner layer forms the<br />

plasmalemma of the ascospore, whilst the outer<br />

membrane forms a perisporic sac. The primary<br />

spore wall is laid down between these two<br />

membranes. It consists of an electron-transparent<br />

endospore and a laminated epispore. The<br />

perisporic sac is variable in thickness and<br />

projects in<strong>to</strong> the epiplasm of the ascus.<br />

At points of contact with the perisporic sac


TUBER (TUBERACEAE)<br />

423<br />

of neighbouring ascospores, the perisporic sac<br />

may be lined by vesicles. Secondary wall formation<br />

results from deposition of material derived<br />

from the epiplasm on<strong>to</strong> the epispore. Dense,<br />

vesicle-bound bodies originally present in the<br />

epiplasm are responsible for the purple coloration<br />

of the secondary spore wall. The pigmented<br />

secondary wall layer is not uniformly thick and<br />

the hyaline striations in the spore wall represent<br />

crevices in the material deposited.<br />

14.5.4 Studies on gene recombination<br />

Ascobolus immersus has proved a useful <strong>to</strong>ol<br />

for interpreting the mechanism of gene<br />

recombination through crossing-over during<br />

meiosis using ascospore colour as a marker.<br />

Recombination is detected by the simple technique<br />

of scoring the ratio of the spores of different<br />

colour in octads of spores shot away from hybrid<br />

apothecia. Although the wild-type strains have<br />

purple ascospores, several series of mutants<br />

with pale spores have been found. When crosses<br />

are made using certain non-allelic spore colour<br />

mutants, recombinants develop resulting from<br />

two types of event: (1) crossing-over, giving<br />

reciprocal recombinants and (2) gene conversion,<br />

yielding non-reciprocal recombinants corresponding<br />

<strong>to</strong> only one of the four products<br />

of meiosis. Gene conversion is a process where<br />

one allele of a gene converts another allele at the<br />

same locus <strong>to</strong> its own type. Crossing-over results<br />

in a 4 : 4 ratio of coloured <strong>to</strong> colourless spores<br />

whilst a conversion is detected by the presence<br />

of coloured and colourless spores in different<br />

ratios e.g. 7 : 1, 6 : 2 or 5 : 3. Conversions have<br />

been interpreted as implying a double-strand<br />

replication of one part of a chromatid whilst<br />

the corresponding part of the other is not<br />

replicated (Lamb, 1996).<br />

14.6 Helvella (Helvellaceae)<br />

Helvella, a genus containing about 40 species, is<br />

mainly distributed in northern temperate areas<br />

(Dissing, 1986; Abbott & Currah, 1997). Helvella<br />

spp. are known as saddle fungi because the<br />

ascocarp is differentiated in<strong>to</strong> a hollow, nonfertile<br />

stipe and a curved, saddle-shaped fertile<br />

part folded over it. They are also called false<br />

morels. Common species are H. crispa (Plate 6f)<br />

and H. lacunosa. Some species form ec<strong>to</strong>mycorrhiza,<br />

and e.g. H. crispa is a mycorrhizal<br />

partner with beech, Fagus sylvatica. A characteristic<br />

feature of the Helvellaceae is that their<br />

ascospores are quadrinucleate. This feature,<br />

also found in certain truffle-like fungi with<br />

hypogeous ascocarps, has been used as evidence<br />

placing genera such as Hydnotrya, species of<br />

which also form sheathing mycorrhizae, in<strong>to</strong><br />

the Pezizales instead of the Tuberales where<br />

they were previously classified (Trappe, 1979).<br />

More recent molecular phylogenetic studies<br />

have confirmed a close relationship between<br />

the Helvellaceae and the Tuberaceae, the main<br />

truffle-containing family (Percudani et al., 1999).<br />

14.7 Tuber (Tuberaceae)<br />

About 100 species of Tuber are known. They<br />

are called the ‘true truffles’, ascomycetes which<br />

form subterranean fruit bodies in which the<br />

hymenium is not open <strong>to</strong> the exterior. It has<br />

been suggested that hypogeous fruiting is an<br />

adaptation for surviving drought and that the<br />

genera of fungi which have adopted this strategy<br />

(including ascomycetes and basidiomycetes)<br />

have evolved from ances<strong>to</strong>rs with epigeous fruit<br />

bodies. According <strong>to</strong> this view the subterranean<br />

ascocarps of Tuber and other ascomyce<strong>to</strong>us<br />

truffle genera are modified apothecia. The term<br />

stereothecium, defined as a more or less solid<br />

fleshy ascoma with asci which are either solitary,<br />

i.e. scattered relatively evenly throughout<br />

the medullary excipulum, or grouped in dispersed<br />

pockets, has been proposed for this<br />

type of ascoma (O’Donnell et al., 1997; Hansen<br />

et al., 2001). Supporting evidence that stereothecia<br />

are modified apothecia is that, during<br />

their on<strong>to</strong>geny, ascocarps are at first open <strong>to</strong><br />

the exterior and close over later as the hymenium<br />

develops (Barry et al., 1993; Callot, 1999;<br />

Janex-Favre & Parguey-Leduc, 2002). This morphological<br />

evidence has been corroborated by


424 HYMENOASCOMYCETES: PEZIZALES (OPERCULATE DISCOMYCETES)<br />

molecular phylogenetic data (Spatafora, 1995;<br />

Percudani et al., 1999; Hansen et al., 2001).<br />

Although Tuber and other ascomycetes were formerly<br />

classified in a separate order (Tuberales),<br />

they are now placed in the Pezizales (Trappe,<br />

1979) as a separate family, the Tuberaceae.<br />

This family is closely related <strong>to</strong> the Helvellaceae<br />

(see above).<br />

14.7.1 The truffle ascocarp<br />

The ascocarp is generally globose, varying in size<br />

from about 1 <strong>to</strong> 8 cm in diameter and, exceptionally,<br />

may weigh up <strong>to</strong> 1000 g. It is differentiated<br />

in<strong>to</strong> an outer, usually dark peridium<br />

which in some species, e.g. T. melanosporum or<br />

T. aestivum, may bear pyramidal scales, and an<br />

inner, fertile gleba. The appearance of the gleba<br />

is marbled because it is traversed by light- and<br />

dark-coloured veins (Fig. 14.7a). The lightcoloured<br />

veins are sterile, consisting of a loose<br />

network of hyphae and air, whilst the darker<br />

veins are fertile, made up of more closely packed<br />

hyphae, paraphyses and asci (Parguey-Leduc<br />

et al., 1991; Barry et al., 1995; Callot, 1999;<br />

Janex-Favre & Parguey-Leduc, 2002).<br />

The asci are unitunicate, subglobose and<br />

contain 2 6 ascospores. They do not discharge<br />

their spores violently, and lack a specialized<br />

apical apparatus or operculum. The ascospores<br />

are at first hyaline, but later develop yellow <strong>to</strong><br />

dark brown (melanized) thick walls which may<br />

be spiny or thrown in<strong>to</strong> reticulate, honeycomblike<br />

folds (Figs. 14.7b, 14.8). Many of the fruit<br />

bodies have a strong smell and flavour, and are<br />

excavated and eaten by animals such as badgers,<br />

wild boar, mice, moles, shrews, squirrels and<br />

rabbits (Trappe & Maser, 1977). Hypogeous<br />

ascocarps (and basidiocarps) form an important<br />

component of their diet. Spore dispersal is<br />

brought about in this way and ascospore germination<br />

is probably enhanced by passage through<br />

the gut of the mammal. Several different volatile<br />

chemical substances have been detected from<br />

truffle fruit bodies, but the most common<br />

and abundant is dimethyl sulphide, attractive<br />

<strong>to</strong> ‘truffle flies’ (Suillia spp.) which lay their eggs<br />

on the ascocarps (Pacioni et al., 1990, 1991).<br />

Similar substances are emitted by stinkhorns<br />

(Phallus spp.), which likewise attract flies (see<br />

p. 590). Claus et al. (1981) have shown that the<br />

steroid hormone 5a-androst-16-en-3a-ol is also<br />

produced by Tuber spp. Since this is the main<br />

sex hormone produced by boars, its presence in<br />

truffles may account for the enthusiasm and<br />

efficiency with which sows locate and excavate<br />

truffles. It is possible that some of the odoriferous<br />

substances are produced by the activity of<br />

microbes associated with ascocarp.<br />

14.7.2 The life cycle of true truffles<br />

Surprise discoveries may happen even in<br />

seemingly well-studied life cycles such as those<br />

of Tuber spp., in which a sympodulosporic conidial<br />

state somewhat resembling Geniculosporium<br />

(see Fig. 12.13) has been described recently (Urban<br />

et al., 2004). It is as yet unclear how frequent<br />

this state is in nature or among other Tuber<br />

spp. and which role, if any, it might play in their<br />

ecology.<br />

The traditional life cycle of Tuber is based<br />

solely on sexual reproduction (see Giovannetti<br />

et al., 1994). The haploid ascospores germinate<br />

<strong>to</strong> form hyphae with monokaryotic segments,<br />

and the mycelium grows <strong>to</strong>wards the roots of<br />

potential mycorrhizal partners, usually trees,<br />

but is unable <strong>to</strong> form mycorrhiza. Anas<strong>to</strong>mosis<br />

between monokaryotic mycelia derived from<br />

different ascospores results in the formation<br />

of a dikaryotic mycelium which forms sheathing<br />

mycorrhiza with suitable hosts (Fasolo-<br />

Bonfante & Brunel, 1972). Ascocarp development<br />

is initiated by the aggregation and differentiation<br />

of hyphae which at this stage remain<br />

attached <strong>to</strong> long roots and obtain nutrients<br />

from the host tree. According <strong>to</strong> Janex-Favre<br />

and Parguey-Leduc (2002), in T. melanosporum<br />

the primordium of the ascocarp consists of an<br />

ascogonium with its trichogyne, surrounded at<br />

the base by an envelope of sterile investing<br />

hyphae. Within the glebal tissues <strong>to</strong> the inside<br />

of the primordium, fertile cells develop. At the<br />

tips of these fertile cells, the two nuclei of the<br />

dikaryon fuse <strong>to</strong> give a diploid nucleus. This<br />

is followed by meiosis and one or more mi<strong>to</strong>ses<br />

so that the ascospores may be uni- or multinucleate<br />

(Delmas, 1978). Ultrastructural studies


TUBER (TUBERACEAE)<br />

425<br />

Fig14.7 Tubermelanosporum,theblackorPe¤rigord truffle. (a) Two fruit bodies, weighing 60 g (left) and 40 g (right).One has been cut<br />

open <strong>to</strong> reveal the black fertile (ascospore-containing) regions interspersed by white (sterile) veins. (b) Ascus containing three<br />

mature ascospores.<br />

Fig14.8 (a c) Tuber rufum. (a) Fruit body in surface view and in section showing the veins. (b) Portion of hymenium. (c) Ascus with<br />

four spiny-walled ascospores. (d,e) Tuber puberulum. (d) V.S. ascocarp showing the structure of the peridium and developing asci.<br />

(e) Mature four-spored ascus showing ascospores with reticulate walls.


426 HYMENOASCOMYCETES: PEZIZALES (OPERCULATE DISCOMYCETES)<br />

have shown that ascosporogenesis resembles<br />

that found in most other Euascomycetes, but<br />

instead of being delimited by the invagination of<br />

an ascus vesicle the ascospores become delimited<br />

individually by vesicles formed by material<br />

derived from lomasomes or from invaginations<br />

of the ascus plasma membrane (Berta & Fusconi,<br />

1983).<br />

At first, the developing ascocarp is attached<br />

by hyphal connections <strong>to</strong> a host tree, i.e. it grows<br />

symbiotically. Eventually it breaks free and<br />

continues <strong>to</strong> develop independently during a<br />

saprotrophic phase of growth. In T. melanosporum,<br />

tufts of hyphae extend in<strong>to</strong> the surrounding soil<br />

from the pyramidal scales on the outside of<br />

the peridium. Application of radioactive tracers<br />

(e.g.<br />

32 PO 4 ,<br />

3 H 2 O and 14 C-labelled mannose) <strong>to</strong><br />

these hyphae is rapidly followed by the appearance<br />

of radioactive material in the inner portions<br />

of the gleba and especially in the fertile<br />

veins (i.e. the dark regions of the ascocarp<br />

interior; see Fig. 14.7a) at rates and in patterns<br />

which could not be accounted for by simple diffusion<br />

(Barry et al., 1994, 1995). This finding<br />

and reports that fruit bodies of truffles may be<br />

found a considerable distance away from living<br />

tree roots suggest that mature ascocarps may<br />

be au<strong>to</strong>nomous and can obtain water and nutrients<br />

directly from the soil, from decaying<br />

roots and faecal deposits from the soil fauna<br />

(Callot, 1999).<br />

The species of Tuber which have been investigated<br />

have a wide mycorrhizal host range,<br />

including Angiosperms and some Gymnosperms.<br />

The host ranges of four species of Tuber are<br />

shown in Table 14.2.<br />

14.7.3 Truffle collecting<br />

The truffles of greatest commercial value are<br />

Tuber melanosporum (the black truffle of Périgord)<br />

and T. magnatum (the white truffle of Piedmont),<br />

which can command prices of up <strong>to</strong> E2000 kg 1<br />

in Continental Europe. Several other species<br />

are also traded. The only common British truffle<br />

which can be used for culinary purposes is<br />

T. aestivum. Périgord and Piedmont truffles are<br />

most abundant in Southern Europe (France,<br />

Spain and Italy) within latitudes 40° and 50°<br />

North in well-drained calcareous soils (Callot,<br />

1999). Both are collected with the aid of pigs<br />

and dogs trained <strong>to</strong> detect them by smell.<br />

The Périgord truffle can also be detected by<br />

the presence of a ‘burnt’ ring-like zone (brûlé)<br />

Table14.2. Host species relationships of four species of Tuber. The main hosts are indicated by black symbols.<br />

Marks in brackets indicate the formation of mycorrhiza with some, but not all members of the host genus<br />

tested. Data summarized from Giovannetti et al. (1994).<br />

Tubermagnatum T.melanosporum T.aestivumgroup T. albidum group<br />

Alnus cordata <br />

Carpinus betulus <br />

Castanea sativa <br />

Cistus (2 spp.) <br />

Corylus avellana <br />

Fagus sylvatica <br />

Ostrya carpinifolia <br />

Populus (2 spp.) <br />

Quercus (6 spp.) () <br />

Salix (2 spp.) <br />

Tilia (3 spp.) <br />

Abies alba <br />

Cedrus (2 spp.) <br />

Pinus (7 spp.) () () ()


MORCHELLA (MORCHELLACEAE)<br />

427<br />

surrounding a tree with roots with mycorrhizal<br />

connections, in which associated herbaceous<br />

plants are wilting or dead. This is partly due<br />

<strong>to</strong> deleterious volatile metabolites from the<br />

Tuber mycelium (Pacioni, 1991) and possibly<br />

also <strong>to</strong> parasitic attack by the mycelium on<br />

roots of herbs. The Périgord truffle is associated<br />

in the wild with the roots of oaks (Quercus spp.)<br />

in France. Truffles are cultivated there in<br />

plantations (truffières) of appropriate species<br />

of oak or on hazel (Corylus avellana). Clonal<br />

material of suitable hazel cultivars may be<br />

used <strong>to</strong> ensure greater yield and uniformity of<br />

cropping (Mamoun & Olivier, 1996). The seedling<br />

roots of potential hosts are inoculated by<br />

dipping them in a suspension of ascospores, or<br />

seedlings can be naturally infected by growing<br />

them close <strong>to</strong> mature mycorrhizal trees where<br />

infection occurs by mycelial contact. Despite<br />

this, truffle yields have fallen continuously, with<br />

about 1000 <strong>to</strong>ns harvested annually in France<br />

around the year 1900 but less than one-tenth<br />

of that yield collected 100 years later (Hall et al.,<br />

2003). A method <strong>to</strong> cultivate truffles on a large<br />

scale under axenic conditions would be a<br />

marvellous achievement, but this is not yet<br />

in sight.<br />

The literature on truffles, stimulated by<br />

their gourmet and high commercial value, is<br />

enormous. Guides <strong>to</strong> identification have been<br />

provided by Gross (1987), Pegler et al. (1993),<br />

Riousset et al. (2001), and a computer-based<br />

interactive key by Zambonelli et al. (2000). More<br />

general accounts of truffle biology and cultivation<br />

have been written by Delmas (1978), Hall<br />

et al. (1994) and Callot (1999).<br />

The ‘desert truffles’, Terfezia spp., are not<br />

closely related <strong>to</strong> Tuber but may instead have<br />

an affinity with the Pezizaceae (Percudani et al.,<br />

1999). Terfezia occurs as a mycorrhizal associate<br />

of shrubs in arid regions of Southern Europe<br />

and the Middle East, where it is consumed as<br />

food and traded on markets. The association<br />

of Terfezia with the roots of shrubs such as<br />

Helianthemum almeriense can greatly improve the<br />

ability of the plant <strong>to</strong> withstand drought stress<br />

and may play an important role in mediterranean<br />

ecosystems (Morte et al., 2000).<br />

14.8 Morchella (Morchellaceae)<br />

The fruit bodies of Morchella spp., the true morels,<br />

are among the most popular and highly prized<br />

edible fungi (Plate 6g). They appear for a few<br />

weeks in spring in cold-temperate regions soon<br />

after snow-melt as the soil becomes warmer and<br />

drier, but they are not confined <strong>to</strong> such areas.<br />

Morchella spp. have two ecological strategies; as<br />

saprotrophic ruderals, fruiting for a relatively<br />

short period (a few years) on disturbed or burnt<br />

ground, or in mycorrhizal association with tree<br />

roots, fruiting over a longer period. The validity<br />

of the alternative lifestyles has been confirmed<br />

by comparative analyses of the iso<strong>to</strong>pes 15 N and<br />

13 C, showing some populations <strong>to</strong> be saprotrophs<br />

whilst others are mycorrhizal (Hobbie et al.,<br />

2001). Opinions on taxonomy vary, with some<br />

authors recognizing about 50 species and others<br />

as few as 3 5 species showing wide phenotypic<br />

variation. Three broad groups of species<br />

have been distinguished, i.e. the half-free morel<br />

(M. semilibera), the black morels (M. elata, M. conica<br />

and M. angusticeps), and the common or yellow<br />

morels (M. esculenta, M. crassipes and M. deliciosa).<br />

Molecular analysis indicates that the black<br />

morels and yellow morels are separate taxonomic<br />

groups (Bunyard et al., 1995; Gessner,<br />

1995). The ascoma of a Morchella consists of<br />

a hollow stipe and a fertile cap thrown in<strong>to</strong><br />

shallow cup-like depressions or alveoli. The<br />

alveoli are lined by asci and paraphyses but<br />

the ridges or ribs which separate the alveoli are<br />

sterile, containing only paraphyses (Janex-Favre<br />

et al., 1998).<br />

The cylindrical, unitunicate, operculate asci<br />

contain eight unicellular ascospores. Karyogamy<br />

occurs prior <strong>to</strong> ascus formation, but croziers<br />

are apparently absent. Following meiosis in<br />

the ascus, there are four successive mi<strong>to</strong>ses so<br />

that the ascospores are multinucleate (Volk &<br />

Leonard, 1990). The tips of the asci are pho<strong>to</strong>tropic<br />

and are directed <strong>to</strong>wards the opening<br />

of the alveolus. The ascospores are often discharged<br />

simultaneously by puffing, generating<br />

air currents which carry clouds of spores well<br />

away from the fruit body (Buller, 1934). They


428 HYMENOASCOMYCETES: PEZIZALES (OPERCULATE DISCOMYCETES)<br />

germinate soon after discharge <strong>to</strong> form a septate<br />

mycelium with multinucleate segments. Frequent<br />

anas<strong>to</strong>mosis may result in the formation<br />

of heterokaryons. Later in the season sclerotia<br />

develop, and it is in this form that the fungus<br />

survives the winter. Ascospores themselves do<br />

not remain viable in the soil for very long<br />

(Schmidt, 1983). The mycelium may also develop<br />

‘muffs’ around the roots of various hosts, mostly<br />

young trees, and within such muffs the mycelium<br />

may penetrate as far as the phloem, an<br />

association which is probably non-mycorrhizal<br />

(Buscot & Roux, 1987; Buscot, 1989). However,<br />

in association with spruce (Picea abies) rootlets<br />

already infected with basidiomyce<strong>to</strong>us mycorrhizal<br />

fungi, the mycelium of Morchella is weakly<br />

mycorrhizal, forming a Hartig net of intercellular<br />

mycelium around one layer of cortical cells of<br />

the host (Buscot & Kottke, 1990). Several morphologically<br />

distinctive types of such secondary<br />

ec<strong>to</strong>mycorrhiza have been observed, often in<br />

association with endobacteria which invade<br />

the hyphae of M. elata making up the Hartig<br />

net and also the host plant cells (Buscot, 1994).<br />

Pure culture synthesis of sheathing mycorrhizae<br />

between Morchella spp. and four species of<br />

Pinaceae has been reported (Dahlstrom et al.,<br />

2000).<br />

The mating system of Morchella is not fully<br />

unders<strong>to</strong>od, and it is as yet uncertain whether<br />

it is homo- or heterothallic. A conidial state,<br />

Costantinella cristata, has been reported <strong>to</strong> develop<br />

following ascospore germination of M. esculenta.<br />

The conidiogenous cells (phialides?) are formed<br />

in verticils arising from lateral branches of the<br />

erect conidiophores. They give rise <strong>to</strong> minute<br />

spherical conidia which do not germinate readily<br />

(Costantin, 1936). Possibly they function as spermatia.<br />

When certain mycelial isolates derived<br />

from single ascospores are confronted in pure<br />

culture, a barrage phenomenon occurs (Hervey<br />

et al., 1978), and this may be due <strong>to</strong> heterokaryon<br />

incompatibility (Volk & Leonard, 1989). The<br />

assumption that sexual reproduction occurs, in<br />

the sense that different parental genomes are<br />

involved in ascocarp formation, is supported<br />

by electrophoretic data confirming that allelic<br />

variation exists in natural populations. It is also<br />

possible that the fertilization events leading <strong>to</strong><br />

the production of different asci within one fruit<br />

body may involve several individuals (Gessner<br />

et al., 1987).<br />

The possibility of cultivating morel ascocarps<br />

commercially is being explored and patents<br />

have been taken out <strong>to</strong> protect the techniques<br />

involved. They are based on the observations by<br />

Ower (1982) that sclerotial development followed<br />

by ascocarp differentiation can be encouraged<br />

by growth of the mycelium on sterilized wheat<br />

grain.


15<br />

Hymenoascomycetes: Helotiales (inoperculate<br />

discomycetes)<br />

15.1 <strong>Introduction</strong><br />

In contrast <strong>to</strong> the Pezizales (see preceding<br />

chapter) which produce apothecia with asci discharging<br />

their spores through a detachable lid at<br />

their apex, the asci of inoperculate discomycetes<br />

liberate their spores either through a valve or<br />

a slit. In the inoperculate as well as operculate<br />

discomycetes, the asci may contain two or more<br />

layers, i.e. they are often described as bitunicate.<br />

However, these layers do not separate during<br />

ascus discharge, i.e. they are non-fissitunicate.<br />

Fine structural details of the asci of Helotiales<br />

have been described by Verkley (1993, 1994,<br />

1996). Two large ecological groups of inoperculate<br />

discomycetes can be distinguished: the<br />

lichenized and non-lichenized species. This feature<br />

correlates approximately with the taxonomy<br />

at the level of orders, and here we shall<br />

discuss the Helotiales (sometimes alternatively<br />

called Leotiales) which contain mostly<br />

non-lichenized fungi. The Lecanorales and other<br />

orders with mainly or exclusively lichen-forming<br />

fungi are described in Chapter 16.<br />

Those relatively few phylogenetic studies that<br />

have so far been performed on the Helotiales<br />

lack the necessary power of resolution <strong>to</strong> delimit<br />

natural groups. Thus, it is not clear at present<br />

whether this order is monophyletic or not,<br />

and several different classification schemes are<br />

in use (Gernandt et al., 2001; Kirk et al., 2001;<br />

Pfister & Kimbrough, 2001). The families<br />

currently associated with the Helotiales are<br />

listed in Table 15.1. Thus circumscribed, the<br />

order Helotiales contains some 2300 species.<br />

The Orbiliaceae, formerly included here (Pfister,<br />

1997), are now considered <strong>to</strong> be more closely<br />

related <strong>to</strong> the Pezizales. Since the conidial forms<br />

of some of them are of interest as nema<strong>to</strong>detrapping<br />

and aquatic fungi, they will be considered<br />

in Chapter 25.<br />

<strong>Fungi</strong> belonging <strong>to</strong> the Helotiales have<br />

adapted <strong>to</strong> several different ecological situations.<br />

Many species are necrotrophic, hemibiotrophic or<br />

biotrophic plant pathogens, and some can cause<br />

considerable damage in economically important<br />

crops. Other species are saprotrophic, colonizing<br />

dead leaves and shoots of herbaceous and woody<br />

plants. Endophytic species are also known, and<br />

it would not be <strong>to</strong>o surprising if many of the<br />

inoperculate discomycetes known as saprotrophs<br />

were found <strong>to</strong> be already present in the living<br />

plant as endophytes. Some species fruit on plant<br />

debris submerged in freshwater streams and have<br />

adapted <strong>to</strong> this habitat by producing conidia of<br />

unusual shapes (Chapter 25). Other species form<br />

ericoid mycorrhizal associations with the roots of<br />

Ericaceae (p. 442). The Thelebolaceae are a group<br />

of coprophilous species.<br />

15.2 Sclerotiniaceae<br />

The Sclerotiniaceae and Rutstroemiaceae are<br />

closely related but can be separated by


430 HYMENOASCOMYCETES: HELOTIALES (INOPERCULATE DISCOMYCETES)<br />

Table 15.1. The families currently placed in or near the Helotiales. Data from Gernandt et al. (2001), Kirk et al.<br />

(2001) and Pfister and Kimbrough (2001). Three families printed in bold are considered in more detail in<br />

this chapter.<br />

Family No. of species Examples<br />

Ascocorticiaceae 3<br />

Bulgariaceae 1 Bulgaria inquinans<br />

Cudoniaceae 10<br />

Cyttariaceae 11 Cyttaria<br />

Dermateaceae (p. 439) 385 Mollisia, Pyrenopeziza, Rhynchosporium,Tapesia<br />

Geoglossaceae 48 Geoglossum,Trichoglossum<br />

Helotiaceae 623 Ascocoryne, Hymenoscyphus, Neobulgaria<br />

Hemiphacidiaceae 12<br />

Hyaloscyphaceae 541 Hyaloscypha, Lachnum (¼ Dasyscyphus)<br />

Leotiaceae 13 Leotia<br />

Loramycetaceae 2<br />

Phacidiaceae 3 Phacidium<br />

Rhytismataceae (p. 440) 219 Rhytisma<br />

Rutstroemiaceae 100 Rutstroemia<br />

Sclerotiniaceae (p. 429) 124 Sclerotinia<br />

Thelebolaceae 15 Thelebolus<br />

Vibrisseaceae 14<br />

DNA-based phylogenetic analyses (Holst-Jensen<br />

et al., 1997). Members of both families produce<br />

stalked apothecia which grow from stromata<br />

located within the colonized host plant tissue.<br />

The apothecia usually develop in spring from<br />

overwintered stromata. The stroma is a food<br />

s<strong>to</strong>rage organ and is usually differentiated in<strong>to</strong><br />

two parts, a rind (cortex) of dark, thick-walled<br />

cells and a medulla of hyaline cells. Two generalized<br />

types of stroma have been distinguished.<br />

To quote from Whetzel (1945),<br />

The sclerotial stroma (commonly called the<br />

sclerotium) has a more or less characteristic form<br />

and a strictly hyphal structure under the natural<br />

conditions of its development. While elements of the<br />

substrate may be embedded in its medulla, they<br />

occur there only incidentally and do not constitute<br />

part of the reserve food supply. The substratal<br />

stroma is of a diffuse or indefinite form, its medulla<br />

being composed of a loose hyphal weft or network<br />

permeating and preserving as a food supply a<br />

portion of the suscept or other substrate (e.g. culture<br />

media).<br />

It is now clear that the sclerotial stroma is typical<br />

of the Sclerotiniaceae where it is often<br />

conspicuous (Plates 7a,b). <strong>Fungi</strong> belonging <strong>to</strong><br />

the Rutstroemiaceae (Plate 7c) produce the less<br />

obvious substratal stroma. The Rutstroemiaceae<br />

grow mainly as saprotrophs, but the Sclerotiniaceae<br />

include some important plant-pathogenic<br />

species. We shall only consider the latter family<br />

further because much more is known about it.<br />

Various types of macroconidia with or without<br />

accompanying microconidia are formed within<br />

the genus Sclerotinia in its widest sense, and species<br />

with different types of conidia are regarded<br />

by many mycologists as belonging <strong>to</strong> distinct<br />

genera. Since these are very closely related,<br />

Sclerotinia is a good example <strong>to</strong> illustrate the flexibility<br />

of asexual reproduction in fungi (Weber &<br />

Webster, 2003). For instance, Sclerotinia fuckeliana<br />

has Botrytis cinerea with polyblastic conidia as its<br />

asexual state (Figs. 15.5a,b) and is thus currently<br />

called Botryotinia fuckeliana. It also produces microconidia<br />

from clustered phialides (Fig. 15.5c).<br />

Sclerotinia (Monilinia) fructigena produces Moniliatype<br />

blas<strong>to</strong>conidia in chains (Fig. 15.3b) and lacks<br />

a microconidial state, whereas S. (Myriosclerotinia)<br />

curreyana produces only Myrioconium-like microconidia<br />

(Figs. 15.1d f) but has no macroconidial


SCLEROTINIACEAE<br />

431<br />

Fig15.1 Sclerotinia<br />

curreyana.(a)T.S.<br />

sclerotium. Note the<br />

stellate pith cells of<br />

the host, Juncus effusus.<br />

(b) T.S. sclerotium<br />

showing<br />

an ascogonium.<br />

(c) Ascus and<br />

ascospores.<br />

(d) Microconidia in<br />

culture. (e) T.S.<br />

spermodochidium on<br />

Juncus effusus.Note<br />

the cavity lined by<br />

phialides.<br />

(f) Microconidia<br />

from host.<br />

state. All of these species can produce apothecia.<br />

This is also true of Sclerotinia sclerotiorum, which<br />

is considered <strong>to</strong> represent Sclerotinia sensu stric<strong>to</strong><br />

(Kohn, 1979). Sclerotium cepivorum produces neither<br />

functional conidia nor apothecia, and the<br />

sclerotia function purely as vegetative propagules,<br />

germinating by hyphal growth. Microconidia<br />

are sometimes produced by germinating<br />

sclerotia, but these do not appear <strong>to</strong> have any<br />

function. The relationship of S. cepivorum with the<br />

Sclerotiniaceae has been deduced from DNAbased<br />

studies (Carbone & Kohn, 1993). Whereas<br />

macroconidia generally germinate readily and<br />

play important roles in the spread of diseases,<br />

microconidia may or may not germinate and<br />

are considered <strong>to</strong> function mainly as spermatia,<br />

i.e. agents of fertilization in sexual reproduction.<br />

15.2.1 Sclerotinia curreyana and S. tuberosa<br />

The apothecia of Sclerotinia (Myriosclerotinia)<br />

curreyana, a pathogen of the rush Juncus effusus,<br />

are common in May. They arise from black sclerotia<br />

in the pith at the base of the Juncus stem<br />

(Plates 7a,b). Infected stems look paler than<br />

healthy stems, and by feeling down <strong>to</strong> the base<br />

of an infected stem the sclerotium can be felt<br />

as a swelling between finger and thumb.<br />

The sclerotium has an outer layer of dark cells<br />

and a pink interior which includes some of the<br />

stellate pith cells of the host (Fig. 15.1a; Plate 7a).<br />

One or several apothecia may grow from a single<br />

sclerotium. The ascospores are released in late<br />

spring and infect the new season’s stems. In<br />

culture, germinated ascospores form a mycelium<br />

which produces microconidia from small phialides<br />

(Fig. 15.1d). Similar clusters of microconidia<br />

can be found on infected Juncus later in<br />

the season (Fig. 15.1f) where they line cavities<br />

beneath the epidermis in the upper part of<br />

infected culms. Whetzel (1946) has used the<br />

term spermodochidium for these microconidial<br />

fructifications (Fig. 15.1e). It is probable that<br />

microconidia play a role in fertilization.<br />

The apothecia of S. (Dumontinia) tuberosa<br />

(Fig. 15.2) are about 2 cm in diameter and arise<br />

from sclerotia within rhizomes of Anemone<br />

nemorosa (Pepin, 1980). They may also occur on<br />

garden Anemone where they are associated with<br />

black rot disease. Microconidia are formed in<br />

culture. Electron microscopy studies of the ascus<br />

wall show that it has a two-layered wall, but the<br />

two layers do not separate from each other,<br />

i.e. the ascus is non-fissitunicate. The ascus apex<br />

contains a thickened dome of wall material with<br />

a central canal. As the ascus explodes, the apical<br />

apparatus is everted (Verkley, 1993).


432 HYMENOASCOMYCETES: HELOTIALES (INOPERCULATE DISCOMYCETES)<br />

Fig15.2 Apothecia of Sclerotinia tuberosa rising from<br />

subterranean sclerotia formed on rhizomes of Anemone<br />

nemorosa.<br />

15.2.2 Monilinia fructigena and M. laxa<br />

Monilinia fructigena is the cause of a brown fruit<br />

rot of apples, pears, plums and other s<strong>to</strong>ne fruits<br />

(Byrde & Willetts, 1977). Although the apothecial<br />

state is only rarely formed, the disease is<br />

common and is transmitted by means of conidia.<br />

Apples and pears showing brown rot bear buffcoloured<br />

pustules of conidia often in concentric<br />

zones (Fig. 15.3a). Sporulation is stimulated by<br />

light, and adjacent zones correspond <strong>to</strong> daily<br />

periods of illumination. The conidia are blas<strong>to</strong>conidia<br />

formed in chains which extend in length<br />

at their apices by budding of the terminal conidium.<br />

Occasionally more than one bud is formed,<br />

and this results in branched chains (Fig. 15.3b).<br />

Conidiogenesis of this type is characteristic<br />

of the anamorphic genus Monilia. Infection of<br />

the fruit is commonly through wounds caused<br />

mechanically or by insects such as codling<br />

moth, wasps or earwigs (Croxall et al., 1951;<br />

Xu & Robinson, 2000). Fruits left lying on the<br />

ground are the source of infection in the following<br />

season. During the winter, infected fruits<br />

become mummified, and the shrivelled fruit<br />

thoroughly colonized by mycelium is interpreted<br />

as the sclerotium. In the following year the<br />

sclerotium may produce further conidial pustules.<br />

Infections can develop as a post-harvest<br />

disease in s<strong>to</strong>red apples, and in some varieties<br />

a twig infection (spur canker) may also occur.<br />

A similar group of diseases of apple and<br />

plum is caused by Monilinia laxa which also has<br />

a Monilia conidial state. In addition <strong>to</strong> fruit rot,<br />

this species causes blossom and shoot blight,<br />

in which infected fresh shoots wilt and become<br />

coated by conidial pustules. Monilinia fructicola<br />

causes brown rot especially of peaches and nectarines<br />

in North and South America, South Africa,<br />

Australia and the Far East, but has not been<br />

reported from Europe. It produces the apothecial<br />

state more readily than the other species (Holtz<br />

et al., 1998), and ascospores released from overwintered<br />

mummified fruits can be a source of<br />

inoculum in the field (Tate & Wood, 2000). These<br />

and a fourth species of the brown fruit rot<br />

complex, M. polystroma, can be distinguished by<br />

means of morphological features and DNA<br />

sequences (van Leeuwen et al., 2002).<br />

15.2.3 Sclerotinia sclerotiorum<br />

This species causes a range of diseases (Sclerotinia<br />

rot, white mould, stalk break) in over 400<br />

cultivated and wild plant species belonging <strong>to</strong><br />

some 75 different families (Boland & Hall, 1994).<br />

The most important crop plants affected are<br />

sunflower, soybean and oilseed rape, with crop<br />

losses approaching 100% under conditions favourable<br />

<strong>to</strong> the disease. Sclerotia are formed on<br />

decaying crop debris and remain viable in a<br />

dormant state in the soil for many years, especially<br />

if deep-ploughed. Sclerotia located within<br />

the <strong>to</strong>p 3 cm of soil germinate <strong>to</strong> produce hyphae<br />

or apothecia in spring. Plants can be infected<br />

either from mycelium or from ascospores; there is<br />

no macroconidial state. Phialidic microconidia<br />

are formed and probably serve as spermatia.<br />

Infection by S. sclerotiorum is often initiated by<br />

a saprotrophic phase on dead leaves or petals<br />

during which mycelial biomass is generated,<br />

prior <strong>to</strong> the attack on the living plant tissue.<br />

Above-ground shoots and, <strong>to</strong> a lesser extent, roots<br />

can be infected, and infections of living tissues<br />

are strongly necrotrophic. This necrotrophic<br />

phase is followed by further saprotrophic<br />

growth and the formation of sclerotia. The<br />

infection cycle in S. sclerotiorum is therefore<br />

tri-phasic. Reviews of the general biology and<br />

pathology of S. sclerotiorum have been written by


SCLEROTINIACEAE<br />

433<br />

Fig15.3 Monilinia fructigena. (a) Apple showing brown rot caused by this fungus, and bearing conidial pustules.The wound serving<br />

as entry point is indicated by an arrow. (b) Blas<strong>to</strong>conidia of the Monilia type.<br />

Purdy (1979), Hegedus and Rimmer (2005) and<br />

Bol<strong>to</strong>n et al. (2006).<br />

Sclerotinia sclerotiorum can infect pollen grains<br />

of crop plants but can also become genuinely<br />

seed-borne, surviving systemically in the embryo<br />

for several years and replacing the rotten tissues<br />

with mycelium and sclerotia (Tu, 1988). Such<br />

seed-borne infections are readily eliminated by<br />

fungicidal seed-dressings, but infections of growing<br />

crops are less easily controlled. Hence,<br />

biological control of S. sclerotiorum has been<br />

attempted. A promising biocontrol agent is the<br />

pycnidial fungus Coniothyrium minitans which<br />

infects and parasitizes hyphae and sclerotia of<br />

S. sclerotiorum in the soil and in plant tissue<br />

(Tribe, 1957; de Vrije et al., 2001). The ability of<br />

C. minitans <strong>to</strong> destroy dormant S. sclerotiorum<br />

sclerotia in the soil is particularly interesting, as<br />

it offers a chance <strong>to</strong> decontaminate infected soil<br />

on which susceptible crop plants could not<br />

otherwise be grown for several years. Gerlagh<br />

et al. (2003) have demonstrated that one or two<br />

conidia of C. minitans are sufficient <strong>to</strong> initiate<br />

infection of a sclerotium. Conidia of C. minitans<br />

can be spread rapidly by the activity of soil<br />

invertebrates, including mites (Williams et al.,<br />

1998). These properties have led <strong>to</strong> the registration<br />

of C. minitans as a commercial biocontrol<br />

agent against S. sclerotiorum. Since C. minitans<br />

colonizing the soil can <strong>to</strong>lerate many fungicides<br />

used against S. sclerotiorum, the integrated<br />

control of Sclerotinia rot is also possible in some<br />

crops (Budge & Whipps, 2001).<br />

Oxalic acid and pH regulation<br />

Like other fungi such as Botrytis cinerea (see<br />

p. 435) and brown-rot basidiomycetes (p. 527),<br />

S. sclerotiorum releases large amounts of oxalic<br />

acid in<strong>to</strong> the infected plant tissue, and this is an<br />

important pathogenicity fac<strong>to</strong>r (Godoy et al.,<br />

1990). Oxalic acid may chelate Ca 2þ ions released<br />

from cell wall degradation, and it also suppresses<br />

the host’s hypersensitive response (Cessna et al.,<br />

2000). Most importantly, however, it acidifies the<br />

infected plant tissue. There is good evidence that<br />

S. sclerotiorum can sense the pH of its environment,<br />

and that it can adjust the production rate<br />

of oxalic acid accordingly. In this way, optimum<br />

conditions are created for the activity of its<br />

pectin-degrading enzymes, especially endopolygalacturonases,<br />

which macerate colonized host<br />

tissues (Rollins & Dickman, 2001). A transcription<br />

fac<strong>to</strong>r encoded by the pac1 gene seems <strong>to</strong> be<br />

involved in regulating the expression of genes<br />

controlled by external pH (Rollins, 2003).<br />

Transgenic sunflower or soybean plants<br />

containing an oxalate oxidase gene from<br />

cereals show good resistance <strong>to</strong> infection by<br />

S. sclerotiorum. This appears <strong>to</strong> be a promising<br />

control strategy for the future, but is currently<br />

still slow in gaining public acceptance (Lu, 2003).<br />

Conventional resistance breeding is also possible,


434 HYMENOASCOMYCETES: HELOTIALES (INOPERCULATE DISCOMYCETES)<br />

although resistance is not usually due <strong>to</strong> major<br />

genes and is only partial. The underlying principle<br />

in runner bean (Phaseolus vulgaris) seems<br />

<strong>to</strong> be an enhanced <strong>to</strong>lerance of oxalic acid or a<br />

restriction of its diffusion through the infected<br />

tissue (Tu, 1985).<br />

Sclerotia<br />

Sclerotia of S. sclerotiorum form readily in culture<br />

and have been the subject of investigations in<strong>to</strong><br />

the physiology of their development (Willetts &<br />

Wong, 1980; Willetts & Bullock, 1992). They are<br />

of the terminal type. General aspects of sclerotial<br />

development have been summarized on<br />

pp. 18 21. The regulation of sclerotium formation<br />

is interesting because it, <strong>to</strong>o, is stimulated<br />

by acid pH, and the Pac1 transcription fac<strong>to</strong>r is<br />

involved (Rollins, 2003). Another signal known <strong>to</strong><br />

be a trigger of sclerotium development is<br />

oxidative stress, e.g. lipid peroxidation or irradiation<br />

with light. Excessive oxidation is prevented<br />

by the synthesis of antioxidants such as b-<br />

carotene or ascorbic acid (vitamin C), and if<br />

high concentrations of these are added <strong>to</strong><br />

cultures of S. sclerotiorum, sclerotium formation<br />

is inhibited (Georgiou & Petropoulou, 2001;<br />

Georgiou et al., 2001).<br />

15.2.4 Sclerotium cepivorum<br />

This fungus causes white rot, the most serious<br />

disease, of Allium spp., especially onions and<br />

garlic. Sclerotia germinate by emitting hyphae<br />

which grow <strong>to</strong>wards the roots of the host plant<br />

and cause necrotrophic infections with maceration<br />

of the root and bulb base tissue. Large<br />

quantities of various pectinolytic enzymes are<br />

secreted (Metcalf & Wilson, 1999). New sclerotia<br />

are formed on and in the decaying bulb tissues.<br />

Sclerotia of S. cepivorum can survive in the soil for<br />

many years or even decades (Coley-Smith, 1959),<br />

and as little as one sclerotium per kg of soil can<br />

cause serious disease losses (Crowe et al., 1980).<br />

Apart from soil fumigation, no effective treatment<br />

of infections or contaminated soil is<br />

available (but see below), and fields may need <strong>to</strong><br />

be abandoned for Allium cultivation once<br />

S. cepivorum has become established.<br />

The root exudates of Allium spp. have long<br />

been known <strong>to</strong> trigger germination of sclerotia<br />

of S. cepivorum. Substances such as alkyl-cysteine<br />

sulphoxides are themselves inactive but are<br />

probably metabolized by soil microbes <strong>to</strong> release<br />

volatile compounds which act as the stimulants<br />

(Coley-Smith & King, 1969; King & Coley-Smith,<br />

1969). One such substance, which is also produced<br />

directly by Allium spp., is diallyl disulphide<br />

(S<strong>to</strong>rsberg et al., 2003). If this is sprayed on<strong>to</strong> an<br />

infested field, it will trigger the germination of<br />

sclerotia which is followed by their death if no<br />

host plants are present. This idea seems <strong>to</strong> hold<br />

potential for the control of S. cepivorum (Coley-<br />

Smith, 1986; Coley-Smith & Parfitt, 1986). Garlic<br />

powder worked in<strong>to</strong> the soil seems <strong>to</strong> have<br />

similar effects, most probably because of the<br />

release of volatile substances from the nonvolatile<br />

water-soluble alkyl cysteine sulphonates,<br />

catalysed by soil bacteria (Fig. 15.4). Biological<br />

control using Trichoderma spp., which secrete<br />

chitinases capable of lysing hyphae<br />

of S. cepivorum, may also be possible (Metcalf &<br />

Wilson, 2001).<br />

15.2.5 The life cycle of Botryotinia<br />

(Sclerotinia) fuckeliana,anamorph<br />

Botrytis cinerea<br />

Because the apothecia of B. fuckeliana are not<br />

commonly seen, the fungus is better known<br />

by its macroconidial state, Botrytis cinerea. This<br />

is ubiqui<strong>to</strong>us on all kinds of moribund plant<br />

material and is also associated with a wide range<br />

of diseases often referred <strong>to</strong> as grey mould.<br />

The name Botrytis cinerea is now known <strong>to</strong> be a<br />

collective name used <strong>to</strong> describe a number of<br />

closely similar, but genetically distinct, strains<br />

(Giraud et al., 1999; Beever & Weeds, 2004). For<br />

this reason some authors prefer <strong>to</strong> write of<br />

a Botrytis of the cinerea type. In-depth treatments<br />

of the biology of Botrytis cinerea have been<br />

compiled by Coley-Smith et al. (1980) and Elad<br />

et al. (2004). In addition <strong>to</strong> B. cinerea, there are<br />

some 20 other Botrytis spp. causing diseases on<br />

a wide range of host plants (Staats et al., 2005).<br />

Macroconidia of B. cinerea are formed on<br />

infected host tissue from dark-coloured branched<br />

conidiophores. The tips of the branches are


SCLEROTINIACEAE<br />

435<br />

Fig15.4 Non-volatile water-soluble<br />

Allium metabolites (left) and their<br />

volatile breakdown products (right).<br />

S-Methyl-L-cysteine sulphoxide (methiin)<br />

is common in many plants, and its<br />

sulphide breakdown products do not<br />

trigger sclerotium germination in<br />

Sclerotium cepivorum. Alliin<br />

(S-2-propenyl-L-cysteine sulphoxide<br />

¼ S-allyl-L-cysteine sulphoxide), isoalliin<br />

and propiin are typical of members of<br />

the genus Allium.Their volatilebreakdown<br />

products, especially mercaptans,<br />

sulphides and disulphides, are potent<br />

triggers of sclerotium germination.<br />

Diallyl disulphide is the major flavour<br />

component of garlic.<br />

thin-walled and bud out <strong>to</strong> form numerous<br />

elliptical multinucleate conidia which are<br />

blas<strong>to</strong>spores (Hughes, 1953). These are easily<br />

detached by the wind, or are thrown off as<br />

the conidiophores twist hygroscopically<br />

(Figs. 15.5a,b). Conidia can also be dispersed by<br />

the fruitfly Drosophila melanogaster (Louis et al.,<br />

1996) and other insect vec<strong>to</strong>rs. Uninucleate<br />

microconidia are formed by clusters of phialides<br />

which arise directly from the mycelium (Weber &<br />

Webster, 2003) or from germinating macroconidia<br />

(Fig. 15.5c). The microconidia have been<br />

claimed <strong>to</strong> be capable of germination (Brierley,<br />

1918) but do not do so in our experience. They<br />

are probably mainly involved in sexual reproduction,<br />

i.e. they function as spermatia. Sclerotia<br />

are formed at the surface of infected tissues and<br />

the fungus overwinters in this form. In spring<br />

the sclerotia may develop <strong>to</strong> give rise <strong>to</strong> tufts<br />

of macroconidia or, much less commonly, <strong>to</strong><br />

apothecia. One or several stalked apothecia may<br />

arise from one sclerotium, with the stalk 1 cm<br />

or more in length and the apothecial disc a few<br />

millimeters in diameter. Botryotinia fuckeliana is<br />

heterothallic with a bipolar mating system. In a<br />

single-ascospore culture, macroconidia, microconidia<br />

and sclerotia can be formed on the same<br />

agar plate (Weber & Webster, 2003), but apothecia<br />

never develop. However, apothecia will form<br />

if microconidia of one mating type are applied<br />

<strong>to</strong> sclerotia of the opposite mating type (Faretra<br />

et al., 1988). Like the great majority of ascomycetes<br />

(Bistis, 1998), B. fuckeliana therefore shows<br />

physiological heterothallism. The life cycle of<br />

this fungus is summarized in Fig. 15.6.<br />

15.2.6 Other life cycles in Sclerotinia<br />

A deviation from the typical ascomycete life<br />

cycle of B. cinerea (Fig. 15.6) is found in Sclerotinia<br />

(Stromatinia) narcissi (Dray<strong>to</strong>n & Groves, 1952).<br />

Of the eight spores formed in the asci of this<br />

fungus, four germinate <strong>to</strong> produce mycelia


436 HYMENOASCOMYCETES: HELOTIALES (INOPERCULATE DISCOMYCETES)<br />

Fig15.5 Botrytiscinerea. (a) Conidiophores developing from a<br />

sclerotium. (b) Apex of conidiophore showing origin of conidia<br />

as blas<strong>to</strong>spores. (c) Conidium germinating <strong>to</strong> produce phialides<br />

and microconidia (after Brierley,1918).<br />

bearing microconidia but no sclerotia, whilst the<br />

other four produce mycelia bearing sclerotial<br />

stromata. Apothecia develop on the strains forming<br />

sclerotia if microconidia are transferred <strong>to</strong><br />

them. Thus the mating behaviour of S. narcissi<br />

differs from that of B. fuckeliana and we can say<br />

that S. narcissi is sexually dimorphic. This type of<br />

behaviour is not common in ascomycetes and it<br />

is possible that such an incompatibility system<br />

has been derived from the more usual system<br />

exemplified by B. fuckeliana by aberrations which<br />

prevent the normal sequence of development of<br />

sexual organs in basically hermaphrodite forms<br />

(Raper, 1959).<br />

A somewhat related phenomenon is found<br />

in Sclerotinia trifoliorum, in which each ascus<br />

contains ascospores of two different sizes. The<br />

four large ascospores germinate <strong>to</strong> give rise <strong>to</strong><br />

homothallic (self-fertile) mycelia, whereas the<br />

four smaller ascospores produce self-sterile<br />

mycelia (Uhm & Fujii, 1983a,b). As discussed earlier<br />

in detail for Saccharomyces cerevisiae (see<br />

Fig. 10.5), the two mating type alleles of ascomycetes<br />

differ strongly in the genes they encode<br />

and are thus termed idiomorphs. Although no<br />

detailed studies seem <strong>to</strong> have been carried out<br />

on the Sclerotiniaceae, in other filamen<strong>to</strong>us<br />

ascomycetes such as Pyrenopeziza brassicae (see<br />

p. 439), heterothallic strains carry either one or<br />

the other of the two idiomorphs whereas in<br />

homothallic species both are fused <strong>to</strong>gether<br />

and expressed simultaneously. In Sclerotinia<br />

trifoliorum, the formation of small ascospores<br />

may be preceded by a unidirectional switch from<br />

homothallic <strong>to</strong> heterothallic, i.e. the deletion of<br />

one of the two mating types during meiosis<br />

(Harring<strong>to</strong>n & McNew, 1997).<br />

Yet another kind of mating behaviour is seen<br />

in Sclerotinia sclerotiorum, which is homothallic.<br />

A single ascospore culture produces microconidia<br />

and sclerotia which bear ascogonial coils<br />

beneath the rind. The transfer of microconidia <strong>to</strong><br />

the sclerotia on the same mycelium results in<br />

the formation of apothecia (Dray<strong>to</strong>n & Groves,<br />

1952). A similar process of self-fertilization also<br />

occurs in Sclerotinia (Botryotinia) porri.<br />

15.2.7 Pathogenicity of Botrytis cinerea<br />

Botrytis cinerea is pathogenic on over 200 species<br />

of plants. Serious diseases of crops are grey<br />

mould of lettuce, <strong>to</strong>ma<strong>to</strong>, strawberry and raspberry,<br />

die-back of gooseberry and damping-off<br />

of conifer seedlings. A special case is bunch rot<br />

of grapes. Under normal circumstances, infected<br />

grapes shrivel and ultimately fall <strong>to</strong> the ground,<br />

forming mummies in which the fungus can<br />

survive the winter. Mummies give rise <strong>to</strong><br />

infective macroconidia in the following spring.<br />

This type of bunch rot causes serious<br />

crop losses in both white and red grapes.<br />

Under certain circumstances and with certain<br />

grape varieties, however, Botrytis causes the<br />

‘noble rot’ in which infections take a milder<br />

course and allow the grape <strong>to</strong> dry out gently,<br />

concentrating its sugar and flavours in the process.<br />

The resulting wine is much sweeter and<br />

richer than normal table wine and is consumed


SCLEROTINIACEAE<br />

437<br />

Fig15.6 Life cycle of Botryotinia fuckeliana (anamorph Botrytis cinerea).Innature,thisfungusoverwintersbymeansofsclerotia<br />

which may germinate in either of two ways.Conidiogenic germination gives rise <strong>to</strong> the macroconidial state which can be formed also<br />

from mycelium.The blastic macroconidia are multinucleate, as are mycelial segments. Phialidic microconidia are formed from<br />

vegetative mycelium or from macroconidia.They are uninucleate, serving mainly as fertilizing agents for sclerotia of opposite mating<br />

type.Fertilization leads <strong>to</strong> carpogenic germination of a sclerotium, i.e. <strong>to</strong> the formation of apothecia, resulting in eight uninucleate<br />

ascospores of either of two mating types.The diploid state is confined <strong>to</strong> the tip of the ascogenous hypha (not shown; see Fig. 8.10).<br />

Open and closed circles represent haploid nuclei of opposite mating type. Key events in the life cycle are plasmogamy (P),<br />

karyogamy (K) and meiosis (M).


438 HYMENOASCOMYCETES: HELOTIALES (INOPERCULATE DISCOMYCETES)<br />

as a dessert wine. Probably the most famous<br />

example is produced in the Sauternes region of<br />

France, using the thin-skinned Sémillon grape<br />

which is particularly susceptible <strong>to</strong> B. cinerea.<br />

Botrytis cinerea can kill its host’s tissue<br />

rapidly and then carries on growing on the<br />

dead remains. It is thus a classical necrotrophic<br />

pathogen. Several research groups have examined<br />

fac<strong>to</strong>rs which may be involved in the<br />

pathogenicity of B. cinerea, but it is as yet<br />

impossible <strong>to</strong> say which ones are the most<br />

important. Quite possibly B. cinerea employs<br />

different strategies for the colonization and<br />

killing of different hosts. This subject has been<br />

reviewed by Prins et al. (2000) and Kars and van<br />

Kan (2004) and is summarized below.<br />

Attachment<br />

Macroconidia of B. cinerea have a hydrophobic<br />

surface, but this is apparently not due <strong>to</strong> the<br />

presence of hydrophobin-type proteins (Doss et al.,<br />

1997). Initial attachment of the macroconidium<br />

<strong>to</strong> the host surface is by weak hydrophobic interactions.<br />

When the germ tube emerges, it secretes<br />

a polysaccharide-based matrix which acts as<br />

a much stronger glue (Doss et al., 1995). This<br />

polysaccharide may be the same as cinerean, a<br />

b-(1,3)-glucan with frequent b-(1,6) cross-linkages<br />

which is produced by B. cinerea from excess<br />

glucose in liquid culture and in infected grapes<br />

(Dubourdieu et al., 1978b; Monschau et al., 1997).<br />

When free glucose becomes scarce, cinerean is<br />

hydrolysed again by extracellular glucanases<br />

(Stahmann et al., 1993). On the plant surface,<br />

the glucan matrix may thus serve in attachment,<br />

as an external carbohydrate reservoir, and as<br />

a matrix for hydrolytic enzymes (Doss, 1999).<br />

Lytic enzymes<br />

After a short period of growth, the germ<br />

tube terminates in a slightly swollen infection<br />

structure which may be considered a rudimentary<br />

appressorium. This is non-melanized, and<br />

thus penetration of the cuticle is probably<br />

mediated mainly by lytic enzymes rather than<br />

turgor pressure (see pp. 381 and 395).<br />

Cutin-degrading enzymes are secreted by<br />

B. cinerea during the initial infection stages<br />

(Comménil et al., 1998), and proteases may also<br />

play a role in pathogenesis (Movahedi & Heale,<br />

1990). Later, a battery of cell wall-degrading<br />

enzymes (especially pectinolytic enzymes) is<br />

produced during the colonization of the host<br />

tissue beyond the initial necrotic lesion. Pectin<br />

seems <strong>to</strong> be a major carbohydrate source for<br />

B. cinerea (Prins et al., 2000). The degradation of<br />

pectin from the middle lamella may also be a<br />

contributing fac<strong>to</strong>r <strong>to</strong> host cell death (Tribe, 1955)<br />

and causes rapid and widespread maceration<br />

of host tissue (Kapat et al., 1998; Kars & van Kan,<br />

2004), which is typical of the necrotrophic<br />

appearance of B. cinerea infections. Oxalic acid is<br />

secreted by B. cinerea as it is by many other fungi,<br />

and its presence is also correlated with tissue<br />

necrosis. However, rather than acting directly as<br />

a <strong>to</strong>xin, it is more likely <strong>to</strong> enhance the activity<br />

of the pectinolytic enzymes which have an acidic<br />

pH optimum, and <strong>to</strong> chelate Ca 2þ ions (Prins et al.,<br />

2000). Substantial quantities of Ca 2þ ions can<br />

be released during pectin degradation from the<br />

carboxylic acid groups of the monomers, galacturonic<br />

acid, which often form calcium salts.<br />

Hypersensitive response<br />

Biotrophic pathogens such as downy or powdery<br />

mildews or rust fungi fail <strong>to</strong> infect incompatible<br />

host plants because these recognize their presence.<br />

One important mechanism of defence is the<br />

hypersensitive response (see pp. 115 and 397) in<br />

which epidermal cells in the vicinity of the<br />

infection site undergo programmed cell death<br />

(Mayer et al., 2001). The hypersensitive response is<br />

accompanied by an ‘oxidative burst’ followed by<br />

the synthesis of phy<strong>to</strong>alexins. With biotrophic<br />

pathogens which require living host cells for<br />

their nutrition, the hypersensitive response is<br />

often sufficient <strong>to</strong> kill the infection unit. If the<br />

necrotrophic B. cinerea attempts <strong>to</strong> infect a host<br />

plant, the hypersensitive response also takes<br />

place, but it fails <strong>to</strong> control the infection because<br />

B. cinerea can exploit the dead cells for nutrition<br />

and initial growth (Govrin & Levine, 2000).<br />

The reactive oxygen intermediates (especially<br />

superoxide and H 2 O 2 ) released during the oxidative<br />

burst may be de<strong>to</strong>xified by the enzymes<br />

superoxide dismutase and catalase, respectively,<br />

which are secreted by B. cinerea and are probably<br />

localized in the glucan matrix surrounding<br />

the infection hypha (Gil-ad et al., 2001).


DERMATEACEAE<br />

439<br />

Further, B. cinerea is known <strong>to</strong> produce laccase<br />

and other enzymes which can degrade or<br />

de<strong>to</strong>xify phy<strong>to</strong>alexins (Prins et al., 2000). ABC<br />

transporters capable of excluding phy<strong>to</strong>alexins<br />

from the hyphal cy<strong>to</strong>plasm have also been<br />

reported from B. cinerea (Schoonbeek et al., 2001;<br />

see also p. 278). Hence, Govrin and Levine (2000)<br />

have suggested that the hypersensitive response<br />

launched by the host actually facilitates, rather<br />

than represses, infection by B. cinerea.<br />

<strong>Fungi</strong>cide resistance<br />

Although biological control strategies against<br />

B. cinerea are being attempted, especially in the<br />

greenhouse and in post-harvest s<strong>to</strong>rage of certain<br />

fruit crops, control in agricultural situations<br />

relies chiefly on the application of fungicides.<br />

This is especially the case for the control of grey<br />

mould on grapevines. Botrytis cinerea has developed<br />

resistance against almost all fungicides in<br />

current use, and this may be due <strong>to</strong> several<br />

fac<strong>to</strong>rs, e.g. the occurrence of sexual reproduction<br />

in the field, the existence of at least two<br />

genetically distinct ‘species’, and the presence<br />

and spread of transposable genetic elements in<br />

one of them (Giraud et al., 1999). All of these<br />

fac<strong>to</strong>rs enhance the genetic diversity of populations<br />

of the pathogen, and thus the chances of<br />

development of fungicide resistance. Mechanisms<br />

of resistance of B. cinerea <strong>to</strong> fungicides<br />

have been discussed by Leroux et al. (2002) and<br />

seem <strong>to</strong> involve strategies also described from<br />

other fungi, i.e. reduced fungicide penetration<br />

in<strong>to</strong> or enhanced export from the hyphae by<br />

means of ABC transporters, enzyme-mediated<br />

de<strong>to</strong>xification and degradation of fungicides,<br />

and mutations leading <strong>to</strong> a reduced binding of<br />

the fungicide <strong>to</strong> its modified target protein.<br />

15.3 Dermateaceae<br />

This family (385 species) is almost certainly<br />

polyphyletic and it will take some time and<br />

numerous further name changes before the<br />

genera are circumscribed <strong>to</strong> the phylogeneticists’<br />

satisfaction. The species included here produce<br />

their apothecia directly on the substratum.<br />

Stromata are absent. The apothecia are small<br />

(less than 1 mm in diameter) and rather inconspicuous,<br />

being coloured in grey, brown or black<br />

<strong>to</strong>nes. The development of apothecia has been<br />

described by Gilles et al. (2001) for Pyrenopeziza<br />

brassicae (Fig. 15.7). Apothecia are formed from<br />

hyphae aggregating in<strong>to</strong> small globular structures<br />

resembling sclerotia or cleis<strong>to</strong>thecia. Later<br />

a pore develops at the apex (Fig. 15.7a), and this<br />

increases in diameter by lateral expansion of the<br />

basal disc (Fig. 15.7b) Meanwhile the asci mature<br />

in the hymenium. Ultimately, a flat apothecium<br />

is formed which possesses a clearly defined<br />

margin typical of the Dermateaceae (Fig. 15.7c).<br />

This developmental pattern has been termed<br />

hemiangiocarpic by Corner (1929). The anamorphs<br />

of Dermateaceae are variable. One very<br />

common form (Cadophora) is Phialophora-like,<br />

i.e. the phialides bear an apical collarette<br />

(Harring<strong>to</strong>n & McNew, 2003). Other forms do<br />

not have phialides, and instead long and transversely<br />

septate conidia are produced more or less<br />

directly from vegetative hyphae.<br />

One large genus (Mollisia) is chiefly saprotrophic<br />

and forms apothecia on dead leaves<br />

and fallen twigs, as exemplified by the ubiqui<strong>to</strong>us<br />

Mollisia cinerea which fruits on dead wood<br />

(Plate 7d). Other members of the family are<br />

hemibiotrophic plant pathogens causing limited<br />

lesions on agricultural crops. Pyrenopeziza<br />

brassicae (anamorph Cylindrosporium concentricum)<br />

causes light leaf spot on winter oilseed rape<br />

(Fig. 15.7) whereas Tapesia yallundae (anamorph<br />

Pseudocercospora herpotrichoides) is the cause of<br />

eyespot at the base of cereal stems, especially<br />

winter wheat, and its sister species, T. acuformis,<br />

causes a similar disease especially on rye. The<br />

conidial Rhynchosporium secalis is the agent of<br />

leaf blotch on a range of cereals. All of these<br />

pathogenic species are phylogenetically closely<br />

related (Goodwin, 2002).<br />

15.3.1 Tapesia yallundae and T. acuformis<br />

Apothecia have been found only recently for<br />

both Tapesia species (see Lucas et al., 2000) and<br />

Pyrenopeziza brassicae (see Gilles et al., 2001). They<br />

have not yet been found for Rhynchosporium<br />

secalis, although the high genetic diversity of<br />

field isolates of this species indicates that<br />

sexual reproduction should occur in nature


440 HYMENOASCOMYCETES: HELOTIALES (INOPERCULATE DISCOMYCETES)<br />

Fig15.7 Development of Pyrenopeziza brassicae ascocarps. (a) Immature apothecium,10 days old on oilseed rape.The apical<br />

pore has just formed. (b) 14-day-old apothecium on oilseed rape.The opening is widening due <strong>to</strong> the expansion of the basal disc.<br />

(c) Mature apothecium 46 days after inoculation on<strong>to</strong> agar. Reprinted from Gilles et al. (2001), with permission from Elsevier.<br />

Images kindly provided by N. Evans.<br />

(Salamati et al., 2000). The mating type idiomorphs<br />

have been characterized for all species<br />

except Tapesia acuformis (Foster & Fitt, 2004), and<br />

they are of the usual heterothallic/bipolar type.<br />

The biology of the two sister species Tapesia<br />

yallundae on wheat and T. acuformis on rye<br />

(formerly called T. yallundae W and R pathotypes,<br />

respectively) is very similar and has been<br />

reviewed by Fitt et al. (1988) and Lucas et al.<br />

(2000). Eyespot is a major disease in winter<br />

cereals growing in cool climates. Infection is<br />

probably mainly by the needle-shaped conidia<br />

which are formed on overwintered stubble and<br />

spread by rain splash. However, ascospores<br />

released from apothecia (Fig. 15.8a) in early<br />

spring are also infectious. If a spore lands on<br />

the coleoptile of a host plant, it germinates and<br />

produces an aggregate of hyphae termed an<br />

infection plaque (Fig. 15.8b). Numerous melanized<br />

appressoria are formed at the interface<br />

of this structure with the host epidermis, so that<br />

infection of susceptible hosts occurs at several<br />

points (Fig. 15.8c). Penetration is probably<br />

mediated by a combination of turgor pressure<br />

and hydrolytic enzymes. The typical eyespot<br />

(Fig. 15.8d) develops as a greyish-brown lesion<br />

around clusters of infection plaques which may<br />

be visible as the ‘pupil’ of the eyespot. Detailed<br />

studies of infection mechanisms have been<br />

published by Daniels et al. (1991, 1995). The presence<br />

of eyespots at the haulm bases renders the<br />

cereal shoots prone <strong>to</strong> collapsing. Further infections<br />

can affect the vascular system, resulting<br />

in poorly developed ‘whiteheads’ containing<br />

inferior grain.<br />

Resistance breeding seems <strong>to</strong> be a promising<br />

strategy for the control of eyespot in cereals<br />

(Lucas et al., 2000). Chemical control is also practised,<br />

but Tapesia spp. have developed resistance<br />

against several types of fungicide (Leroux &<br />

Gredt, 1997).<br />

15.4 Rhytismataceae<br />

The taxonomy of this family is still in a state of<br />

flux (Gernandt et al., 2001). It is sometimes given<br />

ordinal status (Rhytismatales or Phacidiales).<br />

The apothecia are immersed in host tissue or<br />

embedded in a flat stroma. Individual apothecia<br />

become evident when the upper surface breaks<br />

open <strong>to</strong> reveal the hymenium. There are<br />

219 species in this group at present (Kirk et al.,<br />

2001). Most of them are associated with broadleaved<br />

trees or conifers (Cannon & Minter, 1986;<br />

Johns<strong>to</strong>n, 1997). Particularly difficult genera


RHYTISMATACEAE<br />

441<br />

Fig15.8 Infection biology of Tapesia yallundae (a,c,d) and T. acuformis (b). (a) Production of apothecia on overwintered wheat<br />

stubble. (b) SEM of infection plaques on a rye leaf. Runner hyphae (arrows) extend from established plaques on<strong>to</strong> the surrounding<br />

leaf surface. (c) SEM view of a wheat leaf after removal of an infection plaque. Penetration has occurred at numerous points.<br />

(d) Eyespot lesions at the stem bases of wheat plants. (a) kindly provided by P. S. Dyer. (b) and (c) reprinted from Daniels et al.(1991)<br />

with permission from Elsevier; original images of (b d) kindly provided by J. A. Lucas.<br />

in taxonomic terms are Lophodermium and<br />

Lophodermiella which cause needlecast diseases<br />

of Pinus spp. As with many other fungi, there is<br />

a gradient of interactions within the Rhytismataceae,<br />

ranging from the purely endophytic way<br />

of life (Deckert et al., 2001) through saprotrophy<br />

<strong>to</strong> severely pathogenic species. Ortiz-García et al.<br />

(2003) have suggested that at least some<br />

pathogenic species have evolved from endophytic<br />

ances<strong>to</strong>rs.<br />

15.4.1 Rhytisma acerinum<br />

Rhytisma acerinum is common on the leaves of<br />

sycamore, Acer pseudoplatanus, forming black<br />

shiny lesions (tar spots) about 1 2 cm wide<br />

(Fig. 15.9). The lesions arise from infections by


442 HYMENOASCOMYCETES: HELOTIALES (INOPERCULATE DISCOMYCETES)<br />

ascospores released from apothecia on overwintered<br />

leaves. Lesions become visible <strong>to</strong> the naked<br />

eye in June or July, some 2 months after infection,<br />

as yellowish spots which eventually turn<br />

black. Sections of the leaf at this stage show<br />

an extensive mycelium filling the cells of the<br />

mesophyll, and especially the cells of the upper<br />

epidermis. Between the epidermal cells, a conidial<br />

state called Melasmia acerina develops.<br />

This consists of flask-shaped cavities (spermogonia)<br />

which give rise <strong>to</strong> uninucleate curved<br />

club-shaped conidia (spermatia) measuring about<br />

6 1 mm (Figs. 15.10a,b). The spermatia are<br />

exuded from the upper surface of the centre<br />

of the lesion through ostioles in the spermogonial<br />

wall. The spermatia do not germinate, even<br />

on sycamore leaves, and it is believed that they<br />

play a sexual role (Jones, 1925), although this has<br />

not yet been proven. Apothecia begin development<br />

in the portion previously occupied by<br />

spermogonia, and the hymenium is roofed over<br />

by several layers of dark cells formed within the<br />

upper epidermis. The asci complete their development<br />

on the fallen leaves and are ripe about<br />

March <strong>to</strong> April when sycamore leaves of the<br />

new season unfold. The hymenium is exposed by<br />

means of cracks in the surface layer of the fungal<br />

stroma (Fig. 15.9b) and the asci discharge their<br />

spores, sometimes by puffing. Since the ascospores<br />

are very large, their discharge can be<br />

viewed with a dissecting microscope. Although<br />

the ascospores are only projected <strong>to</strong> a height of<br />

about 1 mm above the surface of the stroma, they<br />

are carried by air currents <strong>to</strong> leaves several<br />

metres above the ground. The ascospores are<br />

needle-shaped and have a mucilaginous epispore<br />

which is especially well developed at the upper<br />

end (Fig. 15.10d). This probably helps in attaching<br />

them <strong>to</strong> leaves. Infection occurs by penetration<br />

of the germ tubes through s<strong>to</strong>mata on the<br />

lower epidermis.<br />

Rhytisma acerinum is absent from densely<br />

populated areas, probably because the germination<br />

of ascospores is inhibited by sulphur<br />

dioxide. Greenhalgh and Bevan (1978) have<br />

suggested that the incidence and frequency of<br />

colonization of sycamore leaves by the tar spot<br />

fungus can be used as an accurate visual index of<br />

air pollution, although other interpretations are<br />

possible, such as the removal of fallen leaves<br />

from municipal parks or the drier microclimate<br />

in city centres (Leith & Fowler, 1988).<br />

15.5 Other representatives of the<br />

Helotiales<br />

Many members of the Helotiales are encountered<br />

during fungus forays because they produce<br />

Fig15.9 Rhytisma acerinum. (a) Leaf of sycamore (Acer pseudoplatanus) with developing tar spot lesions. (b) Tarspot from an<br />

overwinteredleaf showing cracking of the surface <strong>to</strong> reveal the hymenia of the apothecia.The flat central zones indicate areas where<br />

spermogonia had been formed during the previous summer.


OTHER REPRESENTATIVES OF THE HELOTIALES<br />

443<br />

Fig15.10 Rhytisma acerinum. (a) T.S. living leaf of Acer pseudoplatanus in June showing spermogonium. (b) Details of cells forming<br />

spermatia. (c) T.S. overwintered leaf of Acer showing the opening of lips of the stromatal surface <strong>to</strong> reveal the apothecial hymenium.<br />

(d) Asci, paraphyses and ascospores. Note the mucilaginous appendage at the upper end of the ascospore.<br />

unusually shaped or brightly coloured fruit<br />

bodies. Good images and keys are given in<br />

Dennis (1981), Breitenbach and Kränzlin (1984)<br />

and Hansen and Knudsen (2000). Very little<br />

is known about Helotiales with small or inconspicuous<br />

apothecia, such as the Hyaloscyphaceae<br />

(Plate 7e).<br />

15.5.1 Geoglossaceae<br />

Trichoglossum is a representative of the Geoglossaceae<br />

(earth-<strong>to</strong>ngues) which form club-shaped<br />

stalked apothecia. Members of this family grow<br />

saprotrophically on the ground, but sometimes<br />

also on dead leaves or amongst Sphagnum<br />

(e.g. Mitrula). An account of the family has been<br />

given by Nannfeldt (1942). Trichoglossum hirsutum<br />

has black, somewhat flattened fruit bodies up<br />

<strong>to</strong> 8 cm high, and grows in pastures and lawns.<br />

The ascospores are long, dark and septate, and<br />

the asci are interspersed by black, thick-walled,<br />

pointed hymenial setae whose function is not<br />

known (Fig. 15.11b). The presence of hymenial<br />

setae separates Trichoglossum from Geoglossum<br />

which grows in similar habitats. The elongated<br />

ascospores of Geoglossum and Trichoglossum are<br />

discharged singly through a minute pore at the<br />

tip of the ascus. When the ascus is ripe, the<br />

pore bursts and one ascospore is squeezed in<strong>to</strong><br />

it, blocking it. The pressure of the ascus sap<br />

behind the spore causes the spore <strong>to</strong> protrude,<br />

at first slowly. When about half the spore is<br />

projecting, the spore gathers velocity and is


444 HYMENOASCOMYCETES: HELOTIALES (INOPERCULATE DISCOMYCETES)<br />

Fig15.11 Trichoglossum hirsutum. (a) Apothecia. (b) Asci,<br />

ascospores, paraphyses and a hymenial seta.<br />

rapidly discharged. Another ascospore immediately<br />

takes the place of the first spore and the<br />

process of discharge is continued until all eight<br />

ascospores have been released in single-file<br />

(Ingold, 1953).<br />

15.5.2 Leotiaceae<br />

Following the separation of the Helotiaceae<br />

(see below), the Leotiaceae now represent only<br />

a small group (13 species) of saprotrophic fungi.<br />

They can be distinguished from the Geoglossaceae<br />

by their brightly coloured ascocarps<br />

and hyaline ascospores (Lizoň et al., 1998).<br />

A well-known example is Leotia lubrica (Plate 7f),<br />

a species colonizing woodland humus and<br />

known colloquially as ‘jelly babies’.<br />

15.5.3 Helotiaceae<br />

Even after the separation of the Leotiaceae, this<br />

is still a very large (4600 spp.) and probably<br />

polyphyletic group. Well-known and widely distributed<br />

saprotrophic genera are Ascocoryne and<br />

Neobulgaria which produce gelatinous pinkish<br />

apothecia on relatively fresh dead wood, or<br />

Chlorociboria with its bright green apothecia<br />

(Plate 7g). Chlorociboria spp. stain the colonized<br />

wood, and this is sometimes used in furniture<br />

making for ornamental inlays. Bisporella citrina<br />

(Plate 7h) is another commonly encountered<br />

species on relatively freshly fallen twigs.<br />

Some members of the Helotiaceae,<br />

notably Hymenoscyphus ericae, as well as some<br />

other ascomycetes belonging <strong>to</strong> the Plec<strong>to</strong>mycetes<br />

(e.g. Pseudogymnoascus, Myxotrichum, Oidiodendron;<br />

see p. 295), can form mycorrhizal<br />

associations with ericaceous plants such as Erica<br />

and Vaccinium. This association is called ericoid<br />

mycorrhiza and has fundamentally different<br />

properties from the vesicular arbuscular<br />

(p. 202), ec<strong>to</strong>mycorrhizal (pp. 525 and 581) and<br />

orchid mycorrhizal types (p. 596). Ericaceous<br />

plants form numerous small lateral roots called<br />

hair roots which consist of a narrow vascular<br />

bundle surrounded by a thin cortex and a thick<br />

epidermal monolayer. When H. ericae infects<br />

individual epidermal cells of its host’s hair<br />

roots, it invaginates the plasmalemma and<br />

forms hyphal coils which superficially resemble<br />

those seen in orchid mycorrhiza (see Fig. 21.2).<br />

Hymenoscyphus ericae is credited with making


OTHER REPRESENTATIVES OF THE HELOTIALES<br />

445<br />

nitrogen and phosphorus available <strong>to</strong> its host<br />

plants which typically grow in situations characterized<br />

by poor soils with acid pH. Good<br />

accounts of ericoid mycorrhiza have been given<br />

by Read (1996), Smith and Read (1997), Berch et al.<br />

(2002) and Peterson et al. (2004).<br />

15.5.4 Bulgariaceae<br />

Bulgaria inquinans forms gelatinous black apothecia<br />

on the bark of recently felled trees (Plate 7i),<br />

especially oak (Quercus), chestnut (Castanea) and<br />

beech (Fagus). Most unusually, the ripe ascus<br />

always seems <strong>to</strong> contain melanized as well as<br />

hyaline spores (Verkley, 1992). This species is<br />

cosmopolitan, and it is possible that it<br />

pre-colonizes the bark of the living tree as an<br />

endophyte.<br />

15.5.5 Cyttariaceae<br />

This family contains some of the most unusual<br />

and striking members of the Helotiales.<br />

Cyttaria spp. live biotrophically on the southern<br />

beech (Nothofagus). Orange-coloured apothecial<br />

stromata which can attain the size of golf balls<br />

arise singly or in clusters from galls on living<br />

tree branches (Plate 7j). Each of the dimples at<br />

the surface of the stroma represents a single<br />

apothecium. The ascospores are dark grey <strong>to</strong><br />

black and continue <strong>to</strong> be discharged in great<br />

numbers even after several days of s<strong>to</strong>rage of<br />

detached stromata in dry conditions. Cyttaria spp.<br />

occur wherever Nothofagus grows, especially in<br />

South America, Australia and New Zealand. The<br />

fruit bodies of some species are edible (Minter<br />

et al., 1987). A review of this enigmatic family of<br />

fungi has been given by Gamundí (1991).


16<br />

Lichenized fungi (chiefly Hymenoascomycetes:<br />

Lecanorales)<br />

16.1 <strong>Introduction</strong><br />

The dual nature of lichens was first hinted at<br />

by de Bary (1866) and clearly recognized by<br />

Schwendener (1867). A lichen is now defined as<br />

a ‘self-supporting association of a fungus (mycobiont)<br />

and a green alga or cyanobacterium<br />

(pho<strong>to</strong>biont)’ (Kirk et al., 2001), ‘resulting in a<br />

stable thallus of specific structure’ (Ahmadjian,<br />

1993). The fungal partner usually contributes<br />

most of the biomass <strong>to</strong> this symbiosis, including<br />

the external surface. It is thus termed the<br />

exhabitant, whereas the unicellular or filamen<strong>to</strong>us<br />

pho<strong>to</strong>biont cells are collectively called the<br />

inhabitant because they are located inside<br />

the lichen thallus (see Ahmadjian, 1993). Most<br />

lichens have a characteristic appearance which<br />

permits their identification if suitable keys are<br />

available (e.g. Purvis et al., 1992; Wirth, 1995a,b;<br />

Brodo et al., 2001). Since the structure of lichens<br />

is almost entirely due <strong>to</strong> the fungal partner,<br />

lichen taxonomy is synonymous with the taxonomy<br />

of the mycobiont.<br />

It is possible <strong>to</strong> grow the algal and fungal<br />

partners of many lichens separately in pure culture<br />

(Ahmadjian, 1993; Crittenden et al., 1995).<br />

Whereas most pho<strong>to</strong>bionts multiply readily in<br />

pure culture, the fungal partner, if it grows at<br />

all, typically shows slow growth as a sterile<br />

leathery mycelium but does not produce the<br />

characteristic lichen thallus. This is in marked<br />

contrast <strong>to</strong> the natural thallus where the<br />

mycobiont displays its full sexual and asexual<br />

cycle, whereas the pho<strong>to</strong>biont cells often appear<br />

swollen and are arrested in their cell cycle, i.e.<br />

their cell division is controlled by the mycobiont.<br />

The nature of the morphogenetic signals<br />

exchanged between the symbiotic partners is as<br />

yet unknown (Honegger, 2001).<br />

Some 13 500 species of lichenized fungi have<br />

been described <strong>to</strong> date. Since lichens are often<br />

conspicuous and have been relatively well<br />

researched over the past 200 years, this number<br />

is not far below the estimated worldwide <strong>to</strong>tal of<br />

some 18 000 species (Sipman & Aptroot, 2001).<br />

In contrast, only about 100 species (40 genera) of<br />

pho<strong>to</strong>bionts are known, although this number<br />

may rise because pho<strong>to</strong>bionts are rarely formally<br />

and fully identified by lichenologists. The most<br />

common pho<strong>to</strong>biont genera are the green algae<br />

Trebouxia (found in about 50% of all lichens)<br />

and Trentepohlia, and the cyanobacterium Nos<strong>to</strong>c.<br />

About 85% of lichenized fungi have a green algal<br />

pho<strong>to</strong>biont, and 10% are associated with a cyanobacterium.<br />

The remaining lichens contain both<br />

a green alga and a cyanobacterium (Honegger,<br />

2001). Most lichenized fungi (498%) belong <strong>to</strong><br />

the Euascomycetes, with only a few imperfect<br />

fungi and some 20 species of Basidiomycota also<br />

entering this type of symbiosis.<br />

Because of their ability <strong>to</strong> <strong>to</strong>lerate repeated<br />

cycles of drying and rehydration and <strong>to</strong> survive<br />

extreme temperatures, high solar irradiation<br />

and other adverse conditions, lichens can colonize<br />

a range of terrestrial habitats not accessible


GENERAL ASPECTS OF LICHEN BIOLOGY<br />

447<br />

<strong>to</strong> higher plants. Lichens are classical pioneer<br />

organisms, e.g. on bare rocks or infertile soils.<br />

Lichens can cause the weathering of rocks by<br />

secreting oxalic acid which reacts chemically<br />

with the rock surface; the rate of degradation<br />

may be 0.5 3.0 mm century 1 (Hale, 1983). Dirina<br />

massiliensis f. sorediata has been shown <strong>to</strong> cause<br />

much more rapid weathering of limes<strong>to</strong>ne surfaces,<br />

including those of his<strong>to</strong>rical monuments,<br />

at a rate of up <strong>to</strong> 2 mm in 12 years (Seaward &<br />

Edwards, 1997). Extensive lichen communities<br />

also exist on the bark and foliage of trees (corticolous<br />

lichens). Additionally, freshwater and<br />

marine species have been described. Lichens<br />

occur in all climatic zones from the Arctic and<br />

Antarctica, where they provide the dominant<br />

vegetation (Seppelt, 1995), <strong>to</strong> the tropics.<br />

Some lichen thalli live for over 1000 years<br />

and can be used for determining the age of rock<br />

surfaces because of their slow growth rate. This<br />

discipline is known as lichenometry (Hale, 1983;<br />

Innes, 1988). It has been applied, for example, <strong>to</strong><br />

date the standing s<strong>to</strong>nes on Easter Island or the<br />

time point of exposure of rock surfaces caused by<br />

avalanches or earthquakes. Crus<strong>to</strong>se lichens are<br />

commonly used for lichenometry because they<br />

have the slowest growth rate. An example is the<br />

‘map lichen’, Rhizocarpon geographicum (Plate 8b;<br />

O’Neal & Schoenenberger, 2003).<br />

A wide range of lichens has been examined<br />

by different research groups. Therefore, in the<br />

present chapter we will give an introduction <strong>to</strong><br />

the general features of lichen biology, followed<br />

by brief profiles of common examples taken from<br />

the Lecanorales, which is by far the largest order<br />

of lichenized fungi. Good general textbooks on<br />

lichens are those by Hale (1983), Ahmadjian<br />

(1993) and Nash (1996a). Richardson (1975) has<br />

written a stimulating account of the importance<br />

of lichens <strong>to</strong> mankind and in natural ecosystems.<br />

16.2 General aspects of lichen<br />

biology<br />

16.2.1 Morphology of the lichen thallus<br />

Lichen thalli come in three basic shapes<br />

crus<strong>to</strong>se (crust-like), fruticose (shaped like a<br />

miniature shrub) or foliose (leaf-like). It should<br />

be noted that these are purely descriptive terms<br />

which have no taxonomic meaning. Intermediate<br />

forms also exist. Good summaries are those by<br />

Büdel and Scheidegger (1996) and Honegger<br />

(2001).<br />

By far the most common type is the crus<strong>to</strong>se<br />

thallus which forms a thin spreading crust<br />

firmly attached <strong>to</strong> the substratum by its entire<br />

lower surface (Plate 8a,b). In the morphologically<br />

simplest crus<strong>to</strong>se lichens, fungal hyphae are<br />

loosely associated with pho<strong>to</strong>biont cells but do<br />

not form a protective upper cortex. Such lichen<br />

thalli appear powdery and are referred <strong>to</strong> as<br />

leprose. They often have a highly hydrophobic<br />

surface. Other crus<strong>to</strong>se lichens produce a thicker<br />

thallus often held <strong>to</strong>gether by mucilage, as in the<br />

gelatinous lichens. In more highly differentiated<br />

crus<strong>to</strong>se thalli, the pho<strong>to</strong>biont cells are positioned<br />

in a defined layer located underneath an<br />

upper cortex formed exclusively by the mycobiont.<br />

The pho<strong>to</strong>biont cells are thus protected<br />

from adverse environmental fac<strong>to</strong>rs. Air spaces<br />

in the pho<strong>to</strong>biont layer and the medulla underneath<br />

permit gas exchange (see Fig. 16.1). The<br />

differentiation of the lichen thallus in<strong>to</strong> horizontal<br />

layers is called stratification. Insquamulose<br />

lichens (Plate 8e), the crus<strong>to</strong>se thallus forms<br />

small scales (squamules) which become partially<br />

raised from the substratum, giving the surface a<br />

scurfy appearance.<br />

The stratification is developed further in the<br />

second thallus type, the foliose lichens (Plate<br />

8c,d) by the development of a lower cortex.<br />

Attachment <strong>to</strong> the substratum is often by bundles<br />

of hyphae termed rhizinae. As a result, the<br />

thallus appears leaf-like or lobed and can be<br />

detached from the substratum without being<br />

damaged.<br />

The third thallus type is called fruticose. Here<br />

the thallus has a shrub-like or branched appearance<br />

and is raised from the substratum (Plate 8f)<br />

or hangs down from it (Plate 8g). In some cases<br />

a fruticose thallus may develop from a basal<br />

crus<strong>to</strong>se or foliose thallus (Plate 8e). Stratification<br />

in fruticose thalli is often tubular/<br />

concentric rather than horizontal, and it resembles<br />

the more complex crus<strong>to</strong>se types in possessing<br />

an outer cortex overlying a pho<strong>to</strong>biont layer


448 LICHENIZED FUNGI (CHIEFLY HYMENOASCOMYCETES: LECANORALES)<br />

Fig16.1 SEM view of a section of<br />

the stratified fruticose thallus of<br />

Anaptychiaciliaris.The globose cells<br />

of the pho<strong>to</strong>biont (Trebouxia)are<br />

located in a loose layer underneath<br />

the thick mucilaginous cortex<br />

produced exclusively by the<br />

mycobiont hyphae. From Bu«del &<br />

Scheidegger (1996), by permission<br />

of Cambridge University Press.<br />

Original image kindly provided by<br />

C. Scheidegger.<br />

and a medulla, but lacking an inner cortex<br />

(Fig. 16.1).<br />

16.2.2 Reproduction of lichens<br />

Sexual reproductive features of lichenized ascomycetes<br />

are largely equivalent <strong>to</strong> those found<br />

in non-lichenized groups. Both homothallic<br />

and heterothallic mating systems are known.<br />

Apothecia with inoperculate asci are especially<br />

common, but perithecia and pseudothecia also<br />

occur. In contrast, no cleis<strong>to</strong>thecia or apothecia<br />

with operculate asci are known among lichenized<br />

fungi. The taxonomic affinities of the<br />

various orders of lichenized ascomycetes are<br />

summarized in Table 16.1. Conidial states, if produced,<br />

are often pycnidial. In most cases, the<br />

pho<strong>to</strong>biont is excluded from the fertile regions<br />

of the mycobiont so that ascospores or conidia<br />

are released without pho<strong>to</strong>biont cells. A new<br />

lichen thallus can be established from fungal<br />

spores only by re-lichenization with suitable<br />

pho<strong>to</strong>biont cells. It is unclear how frequent this<br />

is in nature.<br />

Most lichen thalli therefore produce vegetative<br />

propagules containing both symbionts.<br />

These can be very variable, and an extensive<br />

terminology has evolved <strong>to</strong> describe them (see<br />

Büdel & Scheidegger, 1996). The most common<br />

vegetative propagules are soredia, i.e. small<br />

clumps of hyphae enclosing a few algal cells<br />

(Fig. 16.2). They are produced over the entire<br />

surface of the thallus (Plate 8e; Fig. 16.8) or in<br />

differentiated structures called soralia. Soredia<br />

are usually hydrophobic and are dispersed by<br />

wind, perhaps following their initial detachment<br />

by the impact of a rain drop. Isidia are larger,<br />

upright cylindrical structures which contain<br />

both symbionts. They serve <strong>to</strong> increase the<br />

surface area of the lichen thallus but can also<br />

become detached and then function as vegetative<br />

propagules. In some lichens such as Cladonia,<br />

squamules broken off a vegetative thallus are<br />

capable of establishing a new thallus. Animals<br />

can also play a role in dispersing lichens. Meier<br />

et al. (2002) have shown for Xanthoria parietina<br />

(Plate 8c) that mites feeding on lichen thalli can<br />

spread both the mycobiont and pho<strong>to</strong>biont via<br />

their faecal pellets. Since this lichen does not<br />

produce soredia or isidia, dispersal by invertebrates<br />

could be significant.<br />

16.2.3 Establishment of a lichen thallus<br />

A germinated mycobiont spore in nature may<br />

be able <strong>to</strong> survive for a while as a mycelium<br />

of limited spread, or it can undergo a loose<br />

association with free-living algae not suitable for<br />

an intimate and permanent lichen symbiosis<br />

(Ahmadjian & Jacobs, 1981; Ott, 1987). The initial<br />

stage of the lichen symbiosis is therefore nonspecific.<br />

It results in mycobiont hyphae making<br />

contact with and growing around individual<br />

pho<strong>to</strong>biont cells (Figs. 16.3a,b).


Table 16.1. Summary of the most important orders of lichenized ascomycetes and their characteristic features. Orders have been grouped approximately<br />

according <strong>to</strong> the phylogenetic summary by Grube and Winka (2002).<br />

Order Number of species Features of sexual<br />

reproduction<br />

Lichen thallus Pho<strong>to</strong>biont Taxonomic reference<br />

Lichinales 237 (alllichenized) ascohymenial/apothecial<br />

with pro<strong>to</strong>tunicate asci<br />

crus<strong>to</strong>se, foliose or<br />

fruticose (often<br />

gelatinous)<br />

cyanobacteria Schultz et al.(2001)<br />

Agyriales 98 (mostly lichenized) ascohymenial/apothecial<br />

with bitunicate (nonfissitunicate)<br />

asci<br />

crus<strong>to</strong>se or<br />

squamulose<br />

green algae (with<br />

cyanobacteria in<br />

cephalodia)<br />

Lumbsch et al.(2001)<br />

Gyalectales 108 (alllichenized) ascohymenial/apothecial<br />

with unitunicate asci<br />

mainly crus<strong>to</strong>se mainly green<br />

algae (especially<br />

Trentepohlia)<br />

Lumbsch et al.(2004)<br />

Ostropales<br />

(incl.Graphidales)<br />

1854 (mainly<br />

non-lichenized)<br />

ascohymenial/apothecial<br />

with unitunicate asci<br />

crus<strong>to</strong>se green algae Lumbsch et al.(2004)<br />

Pertusariales 47 (alllichenized) ascohymenial/apothecial<br />

with unitunicate asci<br />

mainly crus<strong>to</strong>se green algae Stenroos and DePriest<br />

(1998)<br />

Lecanorales 7108 (mostly lichenized) ascohymenial/apothecial<br />

with rostrate<br />

(non-fissitunicate) asci<br />

all shapes green algae and/or<br />

cyanobacteria<br />

(see Section16.3)<br />

Pyrenulales 286 (two-thirds<br />

lichenized)<br />

ascohymenial/perithecial<br />

with bitunicate<br />

(fissitunicate) asci<br />

crus<strong>to</strong>se green algae<br />

Verrucariales 720 (mostly lichenized) ascohymenial/perithecial<br />

with bitunicate<br />

(non-fissitunicate) asci<br />

crus<strong>to</strong>se or foliose green algae Wedin et al.(2005)<br />

Arthoniales 1200 (mostly lichenized) ascohymenial/apothecial<br />

with bitunicate<br />

(fissitunicate) asci<br />

mainly crus<strong>to</strong>se green algae Myllys et al.(1998)


450 LICHENIZED FUNGI (CHIEFLY HYMENOASCOMYCETES: LECANORALES)<br />

Fig16.2 Cladonia pyxidata. Soredia containing algal cells<br />

surrounded by fungal hyphae.<br />

Fig16.3 Interactions between pho<strong>to</strong>biontcells (Trebouxia) and mycobiont hyphae. (a,b) Early stages in the establishment of a lichen<br />

thallus in Cladonia cristatella. (a) Single algal cell at point of contact with the mycobiont. (b) Ensheathment of algal cell by the<br />

mycobiont. Both images redrawn and modified from SEM images of Ahmadjian and Jacobs (1981). (c) Formation of tubular<br />

intracellular haus<strong>to</strong>ria by the mycobiont of a simple crus<strong>to</strong>se lichen (e.g. Lecanora conizaeoides).The same pho<strong>to</strong>biont cell may be<br />

penetrated repeatedly. (d) Formation of an intraparietal haus<strong>to</strong>rium by the mycobiont of a stratified lichen. Exchange of nutrients<br />

occurs apoplastically; the walls of both mycobiont and pho<strong>to</strong>biont are surrounded by a common hydrophobin sheath produced by<br />

the former. (c,d) schematic drawings based on the results of Honegger (1986).<br />

If partners are compatible, a pre-thallus<br />

is formed, i.e. a crus<strong>to</strong>se non-stratified cluster<br />

of pho<strong>to</strong>biont cells ensheathed by hyphae of<br />

the mycobiont (Trembley et al., 2002). A prethallus<br />

is also formed by germinating isidia<br />

or soredia. Under suitable conditions, the prethallus<br />

enlarges and stratifies in<strong>to</strong> a mature<br />

thallus; the first sign of this step is the secretion<br />

of mucilage by hyphae at the periphery of the<br />

pre-thallus (Honegger, 1993).<br />

The fac<strong>to</strong>rs determining pho<strong>to</strong>biont<br />

mycobiont specificity and thus permitting the<br />

development of a pre-thallus and mature thallus<br />

are not yet known. In addition <strong>to</strong> genetic determinants,<br />

environmental fac<strong>to</strong>rs must also play<br />

a crucial role as indicated by the difficulties<br />

encountered when trying <strong>to</strong> grow pre-thalli<br />

in<strong>to</strong> fully differentiated thalli in the labora<strong>to</strong>ry<br />

(Ahmadjian, 1993). Recognition is probably a<br />

continuous and multi-step process mediated<br />

by surface molecules such as lectins, and facilitated<br />

by the embedding of both bionts in a<br />

gelatinous matrix of fungal origin (Ahmadjian,<br />

1993).


GENERAL ASPECTS OF LICHEN BIOLOGY<br />

451<br />

A pre-thallus can be formed by fusion of<br />

several genetically distinct isidia, soredia or<br />

pho<strong>to</strong>- and mycobionts, and likewise several<br />

pre-thalli can fuse in the process of thallus<br />

maturation. Thus, a mature thallus may contain<br />

a jigsaw of genetically heterogeneous myco- and<br />

pho<strong>to</strong>bionts (Fahselt, 1996). In other cases, e.g.<br />

the map lichen Rhizocarpon geographicum, the<br />

borders between incompatible thalli are demarcated<br />

by black barrage lines. In some lichens with<br />

a green alga as pho<strong>to</strong>biont, the situation is<br />

further complicated by the inclusion of a second<br />

(cyanobacterial) pho<strong>to</strong>biont. This is then usually<br />

confined <strong>to</strong> discrete regions termed cephalodia<br />

which often differ in their morphology from the<br />

thallus containing the green pho<strong>to</strong>biont (Fig.<br />

16.4).<br />

16.2.4 Lichenicolous fungi<br />

The capture of a compatible pho<strong>to</strong>biont partner<br />

by a germinating fungus spore can be problematic,<br />

especially if the pho<strong>to</strong>biont belongs <strong>to</strong> the<br />

genus Trebouxia which does not seem <strong>to</strong> be<br />

widespread as a free-living organism. One solution<br />

<strong>to</strong> the problem is the recovery of pho<strong>to</strong>bionts<br />

from the propagules of other lichen<br />

species. Certain pho<strong>to</strong>biont strains are favoured<br />

by many taxonomically unrelated mycobionts<br />

(Rikkinen et al., 2002). If these lichens grow<br />

in similar ecological situations, communities are<br />

formed which have a high mycobiont diversity<br />

but share the same or closely related pho<strong>to</strong>biont<br />

strains. Thus, the chances for a germinating<br />

spore <strong>to</strong> salvage compatible pho<strong>to</strong>bionts from<br />

soredia or isidia of other lichen species may be<br />

quite high (Rikkinen, 2003).<br />

Several mycobionts have taken the ultimate<br />

step of poaching their pho<strong>to</strong>biont from an<br />

existing lichen thallus in order <strong>to</strong> establish<br />

their own independent thallus (Ott et al., 1995).<br />

Such organisms are called lichenicolous lichens,<br />

and the phenomenon has been aptly named<br />

‘clep<strong>to</strong>biosis’ (Honegger, 1993). Numerous other<br />

fungi feed on the pho<strong>to</strong>synthetic products of a<br />

lichenized pho<strong>to</strong>biont without ever establishing<br />

an independent thallus, while yet others destructively<br />

parasitize the host lichen (Rambold<br />

& Triebel, 1992; Richardson, 1999; Lawrey &<br />

Diederich, 2003). Such fungi are collectively<br />

called lichenicolous fungi, and they are often<br />

taxonomically related <strong>to</strong> lichenized fungi.<br />

16.2.5 The nutritional basis of the<br />

lichen symbiosis<br />

Hill and Smith (1972) devised a simple and<br />

elegant method termed the ‘inhibition technique’<br />

which has permitted the identification of<br />

carbohydrates secreted by the pho<strong>to</strong>biont. Lichen<br />

thalli are exposed <strong>to</strong> radiolabelled CO 2 , and<br />

after a while an excess of a single unlabelled<br />

Fig16.4 Placopsisgelida, a foliose lichen.The<br />

brightly coloured main thallus containing the<br />

primary (algal) pho<strong>to</strong>biont has produced a<br />

dark central gall-like cephalodium in which<br />

the secondary (cyanobacterial) pho<strong>to</strong>biont is<br />

localized. From Bu«del and Scheidegger (1996),<br />

by permission of Cambridge University Press.<br />

Original image kindly provided by<br />

C. Scheidegger.


452 LICHENIZED FUNGI (CHIEFLY HYMENOASCOMYCETES: LECANORALES)<br />

carbohydrate is added. This will saturate the<br />

uptake system of the mycobiont; only if it is<br />

identical <strong>to</strong> the radiolabelled mobile carbohydrate<br />

exported by the pho<strong>to</strong>biont will the latter<br />

accumulate in the incubation medium, where<br />

its radioactivity can be measured. Such studies<br />

have shown that cyanobacterial pho<strong>to</strong>bionts<br />

export glucose <strong>to</strong> the mycobiont, whereas green<br />

algae export polyols such as erythri<strong>to</strong>l (Trentepohlia),<br />

sorbi<strong>to</strong>l (Hyalococcus, Stichococcus) or ribi<strong>to</strong>l<br />

(Trebouxia, Coccomyxa, Myrmecia) (Ahmadjian,<br />

1993). Tapper (1981) estimated that at least 70%<br />

of the <strong>to</strong>tal pho<strong>to</strong>synthetically fixed carbon is<br />

transferred from the pho<strong>to</strong>biont <strong>to</strong> the mycobiont<br />

in Cladonia convoluta. Once taken up by the<br />

fungus, the transport carbohydrate is rapidly<br />

converted <strong>to</strong> manni<strong>to</strong>l (Lines et al., 1989;<br />

Ahmadjian, 1993).<br />

Cultured algal pho<strong>to</strong>bionts do not secrete<br />

polyols in<strong>to</strong> the medium, and if they are separated<br />

from a fresh lichen thallus, polyol secretion<br />

ceases within a few hours. Nothing appears<br />

<strong>to</strong> be known about the mechanism by which<br />

carbohydrate export from the pho<strong>to</strong>biont is<br />

regulated (Ahmadjian, 1993). Contact of the<br />

mycobiont with its pho<strong>to</strong>biont partner takes<br />

various shapes. The gelatinous extracellular<br />

sheath of cyanobacteria is penetrated by hyphal<br />

protrusions, whereas direct wall-<strong>to</strong>-wall contact<br />

occurs between mycobionts and green algal<br />

pho<strong>to</strong>bionts. Appressorium-like structures presumably<br />

facilitate attachment, and haus<strong>to</strong>ria<br />

may also be formed within the pho<strong>to</strong>biont cells<br />

especially in simple, non-stratified crus<strong>to</strong>se<br />

lichens (Fig. 16.3c). In more highly differentiated<br />

lichens, haus<strong>to</strong>ria are often reduced <strong>to</strong> a pad-like<br />

infection peg appressed <strong>to</strong> but not breaking the<br />

algal wall (Fig. 16.3d). Such structures have been<br />

called intraparietal haus<strong>to</strong>ria (Honegger, 1986).<br />

Fungal hyphae and attached pho<strong>to</strong>biont cells<br />

are often coated by hydrophobin-type proteins<br />

and other hydrophobic molecules secreted by the<br />

mycobiont (Honegger, 1997; Scherrer et al., 2000).<br />

Thus, the main transport route from the pho<strong>to</strong>biont<br />

<strong>to</strong> the mycobiont, in the absence of large<br />

haus<strong>to</strong>rial interfaces, must be through the<br />

apoplast by cell wall contact (Ahmadjian, 1993).<br />

This is different from the elaborate membrane<strong>to</strong>-membrane<br />

contact as found in the haus<strong>to</strong>rial<br />

complexes of arbuscular mycorrhizal fungi (see<br />

Fig. 7.46c) and biotrophic parasites (see Fig. 13.5).<br />

The reason for the reduction in membrane<br />

contact in the lichen symbiosis may lie in the<br />

fact that membranes are among the most easily<br />

damaged structures during the drying and rehydration<br />

cycles <strong>to</strong> which lichens are exposed.<br />

Large haus<strong>to</strong>ria might thus reduce cell viability.<br />

Presumably sufficient carbohydrate leaks out<br />

of the pho<strong>to</strong>biont cells during the frequent<br />

drying rehydration cycles without the need for<br />

intracellular haus<strong>to</strong>ria (Honegger, 1997).<br />

<strong>Fungi</strong> with a cyanobacterium as their primary<br />

or secondary pho<strong>to</strong>biont benefit by receiving<br />

nitrogen in addition <strong>to</strong> carbohydrates. Nitrogen<br />

fixed by cyanobacterial pho<strong>to</strong>bionts is released<br />

<strong>to</strong> the mycobiont as ammonium (NH þ 4 ) and<br />

is incorporated in<strong>to</strong> the amino acid pool as<br />

glutamate (Nash, 1996b).<br />

Whilst the advantages of the lichen symbiosis<br />

<strong>to</strong> the mycobiont are obviously nutritional, there<br />

is no clear evidence of the transfer of any<br />

minerals or other nutrients from the mycobiont<br />

<strong>to</strong> the pho<strong>to</strong>biont. The benefits <strong>to</strong> the pho<strong>to</strong>biont<br />

may include the buffering against adverse<br />

environmental conditions such as high solar<br />

irradiation. The upper cortex of many lichens is<br />

brightly coloured due <strong>to</strong> the presence of<br />

pigments which screen out UV light (see Plate<br />

8b,c). In fact, cortical pigments may filter out as<br />

much as 50% of the incoming light, and this is<br />

particularly important with Trebouxia spp. as<br />

pho<strong>to</strong>biont because these algae favour low light<br />

intensities (Masuch, 1993). Lichen thalli growing<br />

at higher altitudes or on surfaces facing<br />

the sun often contain higher pigment concentrations<br />

than less-exposed thalli. An example of a<br />

light-screen pigment is the polyketide usnic acid<br />

(Fig. 16.5) which is also <strong>to</strong>xic against bacteria,<br />

fungi and other organisms (Elix, 1996; Cocchiet<strong>to</strong><br />

et al., 2002). This substance is produced by several<br />

taxonomically unrelated lichens, but has not yet<br />

been isolated from any non-lichenized fungus.<br />

Another example is the pulvinic acid derivative<br />

vulpinic acid (Fig. 16.5) produced by the wolf’s<br />

lichen, Letharia vulpina. This species is so <strong>to</strong>xic<br />

that its thalli have been used in the past <strong>to</strong><br />

poison foxes and wolves, by laying out animal<br />

carcasses spiked with ground lichen thalli


GENERAL ASPECTS OF LICHEN BIOLOGY<br />

453<br />

Fig16.5 The main routes of secondary metabolism in lichenized fungi.The shikimic acid pathway (<strong>to</strong>p) gives rise <strong>to</strong> vulpinic acid<br />

(¼ pulvinic acid methyl ester) and other pulvinic acid-derived metabolites (see also Fig.19.22).The mevalonic acid pathway gives rise<br />

<strong>to</strong> triterpenes such as sterols (see Fig.13.16) and tetraterpenes such as carotenoids (see Fig. 24.8).The most typical lichen metabolites<br />

are polyketides, especially those synthesized by polymerization of phenolic acids (orsellinic acid) or orcinols derived from them,<br />

giving rise <strong>to</strong> a wide range of depsides (e.g. lecanoric acid) or depsidones (e.g. psoromic acid).This biosynthetic route, although<br />

taking place in the mycobiont, is thought <strong>to</strong> be encouraged by the production of an orsellinic acid decarboxylase inhibi<strong>to</strong>r produced<br />

by the pho<strong>to</strong>biont.Usnic acid is also the product of oxidative coupling of two phenolic-type rings, although the biosynthetic route<br />

does not proceed via orsellinic acid and orcinol.Yet other lichen polyketides (e.g. parietin) arise by cyclization of a single long<br />

polyketide chain; this metabolic pattern is also common in non-lichenized fungi (see Fig.12.46). Modified from Hale (1983),<br />

Masuch(1993)andElix(1996).<br />

(Richardson, 1988; Brodo et al., 2001). Not<br />

surprisingly, lichens containing these and other<br />

<strong>to</strong>xic substances appear <strong>to</strong> be avoided by lichengrazing<br />

animals (Masuch, 1993). Many secondary<br />

metabolites of lichens have acidic properties<br />

and are therefore sometimes collectively called<br />

lichen acids. Most of them are produced by the<br />

mycobiont only in the intact lichen thallus but<br />

not in isolation, indicating that the pho<strong>to</strong>biont<br />

may exert a subtle influence on the secondary<br />

metabolism of the mycobiont. As an example,<br />

Culberson and Ahmadjian (1980) have proposed<br />

that a putative decarboxylase inhibi<strong>to</strong>r secreted<br />

by lichen algae inhibits the conversion of<br />

orsellinic acid <strong>to</strong> phenolic substances which is<br />

common in free-living fungi, leading instead <strong>to</strong><br />

esterification of orsellinic acid or orcinol, and<br />

the accumulation of depside-type lichen acids<br />

(see Fig. 16.5). The <strong>to</strong>pic of secondary metabolism<br />

in lichens is vast and has been reviewed by<br />

Lawrey (1986), Fahselt (1994) and Elix (1996).<br />

The main biosynthetic routes <strong>to</strong>wards secondary<br />

metabolism in lichens are the shikimic acid,<br />

terpenoid and polyketide pathways, of which<br />

the polyketide route is particularly important<br />

(Fig. 16.5).<br />

The benefit conveyed by the mycobiont may<br />

thus be the provision of a ‘pho<strong>to</strong>biont cultivation<br />

chamber’ (Honegger, 2001) which permits the<br />

growth of pho<strong>to</strong>bionts in situations which might


454 LICHENIZED FUNGI (CHIEFLY HYMENOASCOMYCETES: LECANORALES)<br />

be <strong>to</strong>o hostile for free-living forms. The lichen<br />

symbiosis can thus be considered an alternative<br />

adaptation <strong>to</strong> terrestrial life as compared <strong>to</strong><br />

higher plants.<br />

16.2.6 Lichens and pollution<br />

Lichens are particularly sensitive <strong>to</strong> aerial pollutants,<br />

and especially <strong>to</strong> sulphur dioxide, SO 2<br />

(Seaward, 1993; Gries, 1996; Nash & Gries, 2002).<br />

The pho<strong>to</strong>biont appears <strong>to</strong> be generally more<br />

sensitive than the mycobiont. The disappearance<br />

of lichens from the centres of urban and<br />

industrial areas was first recognized by Nylander<br />

(1866), who had already correlated this phenomenon<br />

with aerial pollution. Because different<br />

lichens show a differential sensitivity <strong>to</strong> SO 2 ,<br />

the presence or absence of key species can be<br />

used as an index of the level of air pollution<br />

(Hawksworth & Rose, 1970). The most SO 2 -<br />

<strong>to</strong>lerant lichen, Lecanora conizaeoides, may have<br />

evolved in SO 2 -polluted areas and went on <strong>to</strong><br />

become Northern Europe’s most abundant lichen<br />

by the 1950s (Richardson, 1975). This lichen may<br />

actually require elevated SO 2 levels for good<br />

growth (Nash & Gries, 2002), as shown by its<br />

disappearance from some areas after the implementation<br />

of legislation <strong>to</strong> curb SO 2 emissions.<br />

At the same time, formerly polluted areas are<br />

being re-colonized by many SO 2 -sensitive species<br />

(Rose & Hawksworth, 1981; Seaward, 1993). An<br />

example of this trend has been given by Masuch<br />

(1993) for the city of Munich. Between 1891 and<br />

1956, the ‘lichen desert’ (i.e. lichen-free zone) in<br />

the city centre increased from 8 <strong>to</strong> 56 km 2 , and<br />

then it decreased again, disappearing al<strong>to</strong>gether<br />

by 1983. The size of the ‘lichen desert’ has been<br />

correlated with the degree of SO 2 pollution in<br />

the air. Careful studies of lichen population<br />

dynamics have revealed that lichen species recolonizing<br />

a lichen desert may be different from<br />

those initially present. This phenomenon has<br />

been explained by the eutrophication of urban<br />

habitats, i.e. their enrichment especially with<br />

nitrogen (Seaward, 1997; Seaward & Coppins,<br />

2004). One of these newcomers in urban lichen<br />

deserts is Dirina massiliensis f. sorediata, which is<br />

the cause of a rapid decay of limes<strong>to</strong>ne monuments<br />

(Seaward & Edwards, 1997).<br />

Lichens obtain most of their mineral nutrients<br />

from the air and rainwater in which these<br />

are present only in very low concentrations.<br />

Not surprisingly, therefore, lichens can accumulate<br />

dissolved substances from very dilute<br />

solutions. For instance, lichens concentrate<br />

radioactive nuclides which enter the food chain<br />

lichen!reindeer!man, leading <strong>to</strong> their accumulation<br />

in human tissues (Richardson, 1991).<br />

Lichens are also being used <strong>to</strong> moni<strong>to</strong>r the<br />

radioactive contamination resulting, for example,<br />

from the explosion of the Chernobyl nuclear<br />

reac<strong>to</strong>r in 1986 (Seaward, 2004).<br />

16.2.7 Taxonomy of lichens<br />

The discovery of a fossilized cyanolichen in the<br />

Rhynie chert sediments (Taylor et al., 1997)<br />

indicates that lichens were present some<br />

400 million years ago when the terrestrial habitat<br />

was first colonized. Indeed, there is evidence of<br />

even older lichen-like associations (Yuan et al.,<br />

2005). A huge diversity of lichens exists <strong>to</strong>day,<br />

and there is good phylogenetic evidence that the<br />

lichenized habit has been developed and lost<br />

independently on several occasions in the course<br />

of evolution (Gargas et al., 1995; Lutzoni et al.,<br />

2001). This is also evident from the scattered<br />

placement of lichenized fungi in a wider ascomycete<br />

context. Lutzoni et al. (2001) even suggested<br />

that some of <strong>to</strong>day’s groups consisting<br />

entirely of non-lichenized species, notably the<br />

Plec<strong>to</strong>mycete lineage (Chapter 11), originated<br />

from lichenized ances<strong>to</strong>rs. Part of the proposed<br />

argument is a chemotaxonomic one, i.e. the<br />

presence of numerous secondary metabolites<br />

(especially polyketides) in the lichens and<br />

Plec<strong>to</strong>mycetes, but their absence or less-frequent<br />

occurrence in certain other groups of fungi. As<br />

with many DNA-based analyses, the phylogenetic<br />

arrangement of taxa may vary with the kinds<br />

of sequences used, and other schemes showing a<br />

less scattered distribution of lichenized fungi<br />

within the Ascomycota have been put forward<br />

(Fig. 8.17; Liu & Hall, 2004).<br />

Whereas much work remains <strong>to</strong> be done on<br />

the taxonomy of ascomycetes in general and<br />

lichenized ascomycetes in particular, some<br />

orders are beginning <strong>to</strong> take shape. These are


LECANORALES<br />

455<br />

listed in Table 16.1, summarizing data from<br />

Tehler (1996), Kirk et al. (2001), Ott and Lumbsch<br />

(2001) and Grube and Winka (2002). The orders<br />

Dothideales (Section 17.3), Hypocreales (Section<br />

12.4) and Helotiales (Chapter 15) are not considered<br />

here because they contain only a small<br />

proportion of lichenized species. Some small<br />

families of lichenized fungi (e.g. Baeomycetaceae,<br />

Icmadophilaceae, Umbilicariaceae) are incertae<br />

sedes at the moment, and their accurate placement<br />

will be determined in further studies. Such<br />

studies will have <strong>to</strong> be based on the combined<br />

analysis of several different genes in order <strong>to</strong><br />

obtain a greater degree of confidence in the<br />

resulting phylogenetic trees (Myllys et al., 2002).<br />

Most results so far have been obtained with the<br />

small subunit (18S) ribosomal RNA sequence.<br />

The data summarized in Table 16.1 are <strong>to</strong>o<br />

diffuse <strong>to</strong> be fully unders<strong>to</strong>od at present, but we<br />

note in passing that the occurrence of fissitunicate<br />

asci is not always correlated with ascolocular<br />

development (see p. 459 for an explanation),<br />

i.e. that fissitunicate asci can be found in<br />

perithecia and apothecia, not just pseudothecia<br />

(see Chapter 17). Further, many orders, and<br />

especially the Lecanorales, produce bitunicate<br />

but non-fissitunicate asci, as do some other<br />

Ascomycota, e.g. Helotiales and Pezizales (see<br />

pp. 414. and 429).<br />

16.3 Lecanorales<br />

Members of the Lecanorales produce inoperculate<br />

asci in apothecia. The asci are bitunicate but<br />

non-fissitunicate. The ascus apex is thickened,<br />

and ascospores are discharged when the outermost<br />

wall layer breaks and the innermost layer<br />

protrudes through the pore thus generated, <strong>to</strong><br />

produce an apical beak called rostrum (Figs.<br />

8.12e,f). Rostrate ascus dehiscence is typical of<br />

the Lecanorales.<br />

Over 75% of all lichenized fungi belong <strong>to</strong><br />

this order, making it one of the largest in<br />

the Ascomycota. Most of the best-known and<br />

most readily collected lichens belong <strong>to</strong> the<br />

Lecanorales. Only a few non-lichenized members<br />

of the Lecanorales are known; these are usually<br />

lichenicolous. This order has been divided in<strong>to</strong><br />

42 families (Kirk et al., 2001), of which many still<br />

have an uncertain phylogenetic position and are<br />

currently being circumscribed by DNA sequence<br />

comparisons. Stenroos and DePriest (1998) have<br />

identified five suborders which <strong>to</strong>gether make<br />

up a monophyletic order. We shall consider just<br />

a few representatives <strong>to</strong> indicate the as<strong>to</strong>nishing<br />

morphological and ecological variability of lecanoralean<br />

lichens.<br />

16.3.1 Lecanora<br />

About 300 species of Lecanora have been described,<br />

mainly from temperate climates. Thalli<br />

are usually crus<strong>to</strong>se and are very common on<br />

rock surfaces, including ancient monuments,<br />

dry s<strong>to</strong>ne walls and roof tiles, as well as on the<br />

bark of trees. The pho<strong>to</strong>biont usually belongs <strong>to</strong><br />

the genus Trentepohlia. We have already come<br />

across L. conizaeoides as a particularly SO 2 -<strong>to</strong>lerant<br />

species (p. 454). Because of their exposed habitats,<br />

Lecanora spp. often deposit light-screen<br />

pigments in their upper cortex which give<br />

them a bright yellow coloration. An example is<br />

the xanthone lichexanthone; usnic acid and<br />

pulvinic acid derivatives have also been detected.<br />

The quantity and diversity of pigments may be<br />

greater in thalli exposed <strong>to</strong> higher levels of<br />

irradiation (Obermayer & Poelt, 1992). A common<br />

example of the genus is L. muralis, showing a<br />

typical crus<strong>to</strong>se thallus with apothecia (Plate 8a).<br />

The thalli of Lecanora (Sphaerothallia) esculenta,<br />

a species found from northern Africa <strong>to</strong> western<br />

Central Asia, may roll up and become detached<br />

upon maturity, being blown about by the wind.<br />

They are said <strong>to</strong> be edible. On occasions, windborne<br />

lichen thalli have been so abundant that<br />

the common name ‘manna lichen’ has been<br />

coined for L. esculenta (Richardson, 1988).<br />

16.3.2 Xanthoria<br />

The most abundant species is X. parietina, which<br />

forms bright yellow foliose thalli (Plate 8c) on the<br />

surface of rocks, roofs, trees and farm buildings,<br />

especially near the sea. It is particularly common<br />

in places enriched by manure, e.g. dust from<br />

cattle yards, or from birds. The thallus is lobed<br />

and is attached <strong>to</strong> the substratum by short<br />

rhizinae. The pho<strong>to</strong>biont is the green alga


456 LICHENIZED FUNGI (CHIEFLY HYMENOASCOMYCETES: LECANORALES)<br />

Trebouxia, which forms single globose cells in a<br />

defined layer beneath the upper cortex of the<br />

stratified thallus (Fig. 16.6c). The apothecia are<br />

saucer-shaped and about 2 3 mm in diameter.<br />

They are located on the upper surface of the<br />

thallus, and the algal zone extends in<strong>to</strong> the<br />

apothecial margin (Fig. 16.6a). The ascospores are<br />

at first one-celled, but ingrowth from the wall of<br />

the ascospore eventually divides the contents<br />

of the spore in<strong>to</strong> two. The yellow colour of the<br />

thallus is due <strong>to</strong> the presence of the anthraquinone<br />

parietin (Fig. 16.5) in the upper cortex.<br />

Unusual carotenoids are also produced by<br />

Xanthoria spp. (Czeczuga, 1983). Xanthoria parietina<br />

does not produce soredia, but it is very<br />

abundant none the less. One reason for this may<br />

be that germinating ascospores display a tendency<br />

<strong>to</strong>wards clep<strong>to</strong>biosis, i.e. the theft of<br />

pho<strong>to</strong>biont cells from soredia or mature thalli<br />

of other lichens (Ott, 1987). Another means of<br />

dispersal may be by mite browsing and feeding<br />

as mentioned earlier.<br />

16.3.3 Peltigera<br />

About 45 60 species are known, and the genus<br />

has been thoroughly examined by Miadlikowska<br />

and Lutzoni (2004). They and many other workers<br />

now consider it <strong>to</strong> be part of a separate order,<br />

Peltigerales, which is closely related <strong>to</strong> Lecanorales.<br />

Species of Peltigera form large lobed leaf-like<br />

thalli attached <strong>to</strong> the ground or <strong>to</strong> rocks by<br />

groups of white rhizinae. The thallus is rather<br />

fleshy and is highly stratified (Fig. 16.7a). The<br />

commonest species are P. polydactyla and P. canina<br />

(Plate 8d), both of which have now been split up<br />

in<strong>to</strong> several species. They grow among grass on<br />

heaths, on sand dunes and on rocks amongst<br />

moss. The usual pho<strong>to</strong>biont is the cyanobacterium<br />

Nos<strong>to</strong>c. In some species, however, the<br />

primary pho<strong>to</strong>biont can be either a Nos<strong>to</strong>c,<br />

giving rise <strong>to</strong> the usual greyish-black thallus, or<br />

a green alga (Coccomyxa), in which case the<br />

thallus is vividly green, with Nos<strong>to</strong>c sometimes<br />

present as a secondary pho<strong>to</strong>biont in cephalodia<br />

(Brodo & Richardson, 1978). The two different<br />

Fig16.6 Xanthoria parietina. (a) V.S. thallus and apothecium showing the extension of the algal zone in<strong>to</strong> the apothecium. (b) Asci,<br />

paraphyses and two germinating ascospores. (c) V.S. thallus.


LECANORALES<br />

457<br />

Fig16.8 Cladonia pyxidata. Primary squamulose thallus<br />

bearing funnel-shaped podetia. Note the granular soredia<br />

outside and inside the podetia.<br />

contain algal cells. The red colour of the<br />

apothecia is due <strong>to</strong> pigments in the tips of the<br />

paraphyses (Fig. 16.7b).<br />

Fig16.7 Peltigera polydactyla. (a) V.S. thallus. (b) Ascus,<br />

ascospores and paraphyses.<br />

forms of the same lichen are then termed a<br />

morphotype pair or ‘lichen chimera’. The<br />

apothecia of Peltigera are reddish-brown, folded<br />

extensions of the thallus (Plate 8d) which do not<br />

16.3.4 Cladonia<br />

There are about 400 species of Cladonia, some of<br />

them extremely common, growing in heaths,<br />

moors and elsewhere on rocks and walls. There<br />

are two kinds of thallus. The primary thallus is<br />

squamulose, and the secondary thallus is upright<br />

and cylindrical (fruticose), often consisting of a<br />

hollow stalk which bears the apothecium at its<br />

tip. Such an apothecium-bearing vertical thallus<br />

arising from a horizontal primary thallus is<br />

called a podetium and is typical of the genus<br />

Cladonia. InC. pyxidata, the podetium opens out<br />

in<strong>to</strong> a cup (Fig. 16.8), and the apothecia ultimately<br />

develop at the rim of the podetium. In C.<br />

floerkiana (Plate 8e), the podetia are shrub-like<br />

and bear their red apothecia as terminal heads.<br />

The colloquial name for this species is ‘British<br />

soldiers’. The podetia frequently bear the granular<br />

soredia which contain algal cells surrounded<br />

by fungal hyphae (Fig. 16.2). In windtunnel<br />

experiments using C. pyxidata, Brodie and<br />

Gregory (1953) showed that soredia were blown<br />

away from the funnel-shaped podetia at wind<br />

speeds as low as 5.4 7.2 km h 1 although they<br />

were not removed from horizontal glass slides at<br />

the same wind speeds. They suggested that<br />

funnel-shaped structures generate eddy currents


458 LICHENIZED FUNGI (CHIEFLY HYMENOASCOMYCETES: LECANORALES)<br />

when placed in a windstream and that eddy<br />

currents effectively remove soredia.<br />

Usnic acids (Fig. 16.5) are particularly common<br />

in Cladonia spp., and their concentration in the<br />

lichen thallus has been shown <strong>to</strong> increase linearly<br />

with the intensity of UV light, supporting a<br />

possible role as a light screen (Rundel, 1969).<br />

Further, sun-exposed thalli of Cladonia spp. appear<br />

yellowish whereas those in more shaded habitats<br />

are greyish-green, although the pigment responsible<br />

is not usnic acid.<br />

The reindeer lichen genus Cladina is closely<br />

related <strong>to</strong> Cladonia (Stenroos & DePriest, 1998).<br />

The most common species are C. rangiferina (Plate<br />

8f) and C. stellaris, which are a major winter food<br />

for grazing animals such as reindeer or caribou<br />

in northern boreal forests (Richardson, 1988).<br />

These lichens provide an important component<br />

of the ground cover grazed by animals, and are<br />

also used by Laplanders <strong>to</strong> make hay for their<br />

animals. Reindeer lichens are popular in<br />

Germany as decorations on wreaths and are<br />

also well-known among model railway enthusiasts<br />

and architects who use the highly<br />

branched fruticose thalli as miniature trees<br />

(Kauppi, 1979; Richardson, 1988).


17<br />

Loculoascomycetes<br />

17.1 <strong>Introduction</strong><br />

The characteristic feature of this group is that<br />

the ascus is bitunicate and fissitunicate; it has<br />

two separable walls (see p. 240). The outer wall<br />

(ec<strong>to</strong>tunica or ec<strong>to</strong>ascus) does not stretch readily,<br />

but ruptures laterally or at its apex <strong>to</strong> allow<br />

the stretching of the thinner inner layer, the<br />

endotunica or endoascus (Figs. 17.1a c). Asci<br />

are generally non-amyloid. The fruit body with<br />

asci is regarded as an ascostroma, and each<br />

cavity in which asci develop is termed a locule.<br />

In contrast <strong>to</strong> the Hymenoascomycetes, in which<br />

ascocarps develop following plasmogamy and<br />

the pairing up of two genetically dissimilar<br />

nuclei (ascohymenial development), in the<br />

Loculoascomycetes the ascoma is already present<br />

before the compatible nuclei are brought <strong>to</strong>gether<br />

(Barr & Huhndorf, 2001). The development of<br />

asci in pre-formed locules is called ascolocular.<br />

The ascostroma has therefore been defined as<br />

an aggregation of vegetative hyphae not resulting<br />

from a sexual stimulus (Wehmeyer, 1926).<br />

However, Holm (1959) has questioned the accuracy<br />

of this definition, since examples are known<br />

where the ascocarps do develop following the<br />

pairing of nuclei (Shoemaker, 1955). Within the<br />

developing ascocarp, one or more locules are<br />

formed by the downgrowth of pseudoparaphyses<br />

(see below) and the development of asci. One or<br />

more ostioles then develop by the breakdown<br />

(lysis) of a pre-formed mass of tissue (lysigenous<br />

development). Where a single locule develops, a<br />

structure resembling a perithecium results and,<br />

although this term is commonly used for such<br />

loculoascomycete fruit bodies, they should<br />

strictly be called pseudothecia (see p. 245).<br />

Although a mature ascostroma with several<br />

locules can be superficially similar <strong>to</strong> the perithecial<br />

stroma of Pyrenomycetes such as<br />

Hypoxylon (Fig. 12.11) or Claviceps (Fig. 12.27), the<br />

difference is that in the pyrenomycete stroma<br />

each fertile region (perithecium) is surrounded<br />

by a wall whereas the locules of the ascostroma<br />

are not (Alexopoulos et al., 1996). However, the<br />

ascocarp of Loculoascomycetes does not always<br />

take the form of a pseudothecium. In some groups<br />

it may be an apothecium, a hysterothecium (an<br />

elongate ascoma with a slit-like opening) or a<br />

cleis<strong>to</strong>thecium (Barr & Huhndorf, 2001). These<br />

deviations from the classical pseudothecium are<br />

found for example in lichenized Loculoascomycetes<br />

(see Table 16.1).<br />

The name Loculoascomycetes was coined<br />

by Luttrell (1955) and corresponds <strong>to</strong> the<br />

Ascoloculares of Nannfeldt (1932). The group<br />

has also been named the Dothideomycetidae (see<br />

Kirk et al., 2001). It is very large, with about 900<br />

genera and over 7000 species. Most members are<br />

terrestrial, growing as saprotrophs, endophytes<br />

or parasites on the shoots and leaves of herbaceous<br />

or woody plants and may cause diseases of<br />

economic significance, but some grow in freshwater<br />

or the sea and others in dung or soil.<br />

There is a very wide range of anamorphs, some<br />

hyphomyce<strong>to</strong>us, others pycnidial (Sivanesan,<br />

1984; Seifert & Gams, 2001).<br />

It is most unlikely that the Loculoascomycetes,<br />

as currently classified, are monophyletic


460 LOCULOASCOMYCETES<br />

Fig17.1 Sexualreproductionin Pleosporascirpicola. (a) Young ascus withundifferentiated cy<strong>to</strong>plasm.Note the thick ascus wall.<br />

(b)Maturingasciwithcy<strong>to</strong>plasmcleavedin<strong>to</strong>ascospores.(c)Matureascuswithmuriformascospores.Theascusis shownafewseconds<br />

before discharge; the ec<strong>to</strong>tunica hasruptured (arrow), and themore flexible endotunica hasbeenforced <strong>to</strong> expandby theincreasing<br />

turgor pressure. (d) Discharged ascospore mountedin Indian ink <strong>to</strong> show the mucilaginous extensions. Allimages <strong>to</strong> same scale.<br />

(Berbee, 1996; Dong et al., 1998; Silva-Hanlin &<br />

Hanlin, 1999). Molecular analyses based on a relatively<br />

small number of representatives of ascomycete<br />

groups suggest that the most closely related<br />

groups are Pyrenomycetes (perithecioid fungi)<br />

belonging <strong>to</strong> Diaporthales, Hypocreales, Microascales<br />

and Sordariales (see Chapter 12; Spatafora,<br />

1995). Several orders are currently included in<br />

the Loculoascomycetes (Barr & Huhndorf, 2001;<br />

Kirk et al., 2001) but we shall consider representatives<br />

only of the two most important, the<br />

Pleosporales and Dothideales. Because the<br />

arrangement of orders and especially families<br />

is still in a state of flux, we shall not discuss it.<br />

17.2 Pleosporales<br />

This is a large group of ascomycetes believed <strong>to</strong><br />

be monophyletic on the basis of molecular<br />

evidence (Berbee, 1996). Several economically<br />

important genera of plant pathogens are<br />

included, such as Cochliobolus, Phaeosphaeria and<br />

Pyrenophora, parasitic on grasses and cereals.<br />

Pleospora, Lewia and Lep<strong>to</strong>sphaeria are common<br />

endophytes, saprotrophs or parasites of other<br />

herbaceous plants. They and their anamorphs<br />

may also have significance as human allergens,<br />

as human pathogens and in the production<br />

of myco<strong>to</strong>xins. Sporormiella fruits on herbivore<br />

dung.<br />

The development of pseudothecia in these<br />

forms conforms <strong>to</strong> the Pleospora type of Luttrell<br />

(1951). Ascogonia arise within a stroma and,<br />

in the region of the ascogonia, a group of<br />

vertically arranged septate hyphae appears,<br />

each hypha arising as an outgrowth from a<br />

stromatal cell. These hyphae are capable of<br />

elongating by intercalary growth and are<br />

termed pseudoparaphyses (Luttrell, 1965).


PLEOSPORALES<br />

461<br />

Pseudoparaphyses arise near the upper end of<br />

the cavity and grow downwards. Their tips soon<br />

intertwine and push between the other cells of<br />

the stroma so that free ends are seldom found.<br />

They may thus be distinguished from the true<br />

paraphyses of other fungi (e.g. Sordaria) which are<br />

formed from hyphae attached at the base of the<br />

cavity, extend upwards and are free at their<br />

upper ends. They may also be distinguished from<br />

apical paraphyses which are attached above,<br />

arising from a clearly defined meristem near<br />

the apex of a perithecium, and form a welldefined<br />

palisade of hyphae free at their lower<br />

ends (see the Nectria type of development, p. 337).<br />

In the Pleospora type of development, asci arise<br />

amongst the pseudoparaphyses at the base of the<br />

cavity and grow up between them. The ostiole<br />

develops lysigenously, i.e. by breakdown of preexisting<br />

cells. Development of this general type<br />

has been described in P. herbarum (Wehmeyer,<br />

1955; Corlett, 1973), Lep<strong>to</strong>sphaeria (Dodge, 1937),<br />

Sporormiella (Arnold, 1928) and other fungi (see<br />

Luttrell, 1951, 1973). A more recent discussion<br />

of pseudoparaphyses development in the Pleosporales<br />

and its taxonomic implications has been<br />

written by Liew et al. (2000).<br />

17.2.1 Lep<strong>to</strong>sphaeria<br />

Lep<strong>to</strong>sphaeria species fruit on moribund leaves<br />

and stems of herbaceous plants. There are<br />

probably some 100 species, many growing on a<br />

wide range of hosts, but others are confined <strong>to</strong><br />

one host plant. Although most are saprotrophic<br />

or only weakly pathogenic, some are troublesome<br />

pathogens, e.g. L. coniothyrium, the cause of<br />

cane blight of raspberry, and L. maculans which<br />

causes blackleg of oilseed rape and other<br />

brassicas. Characteristic features are the fusoid,<br />

yellow or pale brown ascospores with two or<br />

more transverse septa. Anamorphic states are<br />

pycnidial (see Table 17.1).<br />

Lep<strong>to</strong>sphaeria acuta fruits in abundance in<br />

spring at the base of overwintered, decorticated<br />

stems of stinging nettles (Urtica dioica). The black<br />

shining pseudothecia are somewhat conical and<br />

flattened at the base (Fig. 17.2a). Bitunicate asci<br />

elongate within a pre-existing group of branching<br />

pseudoparaphyses, and close examination of<br />

the direction of growth and branching indicates<br />

that the pseudoparaphyses may be ascending<br />

and descending. The ostiole of the perithecium<br />

is formed lysigenously by breakdown of a preexisting<br />

mass of thin-walled cells (Fig. 17.3a).<br />

The bitunicate structure of mature asci is<br />

difficult <strong>to</strong> discern because, as the ascus expands,<br />

the inner wall protrudes through a thin area<br />

in the outer wall at the ascus tip (Fig. 17.3e)<br />

and then the inner wall extends. Thus the<br />

ascus tip in expanded asci is single-walled.<br />

The ascospores have about 11 transverse septa<br />

and are discharged successively at intervals of<br />

about 5 s.<br />

Associated with the thick-walled conical<br />

pseudothecia on the nettle stem are thinnerwalled,<br />

slightly smaller, globose pycnidia with<br />

cylindrical necks (Figs. 17.2b and 17.3b). The<br />

cavity of the pycnidia is lined by small spherical<br />

cells which give rise <strong>to</strong> numerous rod-shaped<br />

conidia (Fig. 17.3c). These are dispersed by rain<br />

splash or in water films and are capable of<br />

germination, which suggests that they do not<br />

function as spermatia. Such pycnidia have been<br />

Table 17.1. Anamorphic states of some species of Lep<strong>to</strong>sphaeria and Phaeosphaeria.<br />

Teleomorph Anamorph Disease<br />

L. bicolor Stagonospora sp. Sugarcane leaf scorch<br />

L. coniothyrium Coniothyrium fuckelii Raspberry cane blight<br />

L. maculans Phoma lingam Blackleg of oilseed rape and other brassicas<br />

P. avenaria Stagonospora avenae Oatleaf blotch<br />

P. microscopica Phaeosep<strong>to</strong>ria festucae Leaf spot of fescue and other grasses<br />

P. nodorum Stagonospora nodorum Glume and leaf blotch of wheat and barley


462 LOCULOASCOMYCETES<br />

Fig17.2 Lep<strong>to</strong>sphaeriaacuta on overwinterednettle stem.<br />

(a) Pseudothecia. (b) Phoma anamorph.Bothimages <strong>to</strong> same<br />

scale.<br />

named Phoma acuta, and culture studies have<br />

confirmed that this stage is the conidial state of<br />

L. acuta (Müller & Tomasevic, 1957). Pseudothecia<br />

of L. acuta ripen in the spring and discharge<br />

ascospores as the new season’s nettle shoots are<br />

elongating. There are no obvious disease symp<strong>to</strong>ms<br />

on infected plants during the summer and<br />

the fungus persists as a symp<strong>to</strong>mless endophyte,<br />

with the pseudothecia and pycnidia developing<br />

during the winter on dead stems of the past<br />

season’s growth.<br />

17.2.2 Lep<strong>to</strong>sphaeria and Phaeosphaeria<br />

An unusually high degree of taxonomic confusion<br />

has arisen in attempts <strong>to</strong> delimit and define<br />

several important plant-pathogenic species<br />

which clearly belong <strong>to</strong> the Pleosporales but<br />

show immense variations in their anamorph<br />

and teleomorph features. One example is a group<br />

of coelomyce<strong>to</strong>us species comprising Sep<strong>to</strong>ria<br />

with multiseptate conidia more than 10 times<br />

longer than wide, Stagonospora with multiseptate<br />

conidia less than 10 times longer than wide<br />

(Fig. 17.4), and Phoma with aseptate globose<br />

or slightly elongated conidia (Figs. 17.6 and<br />

17.7). The most common sexual states associated<br />

with Sep<strong>to</strong>ria and Stagonospora are Lep<strong>to</strong>sphaeria<br />

and Phaeosphaeria (Leuchtmann, 1984). As with<br />

the anamorphs, the morphological features of<br />

these teleomorphs show transitions, but a broad<br />

distinction can be made by Lep<strong>to</strong>sphaeria being<br />

associated mainly with dicotyledons whereas<br />

Phaeosphaeria is pathogenic on monocotyledons<br />

(Cunfer & Ueng, 1999). These two genera have<br />

also been separated by critical DNA sequence<br />

analyses (Câmara et al., 2002).<br />

Among the cereal pathogens belonging here,<br />

Phaeosphaeria (Stagonospora) nodorum has two<br />

formae speciales, one on wheat and the other on<br />

barley, whereas Phaeosphaeria avenaria<br />

(Stagonospora avenae) is a more diverse species<br />

complex (Ueng et al., 1998) but is not as serious a<br />

pathogen in most agricultural situations. These<br />

fungi cause leaf blotch diseases which are very<br />

common in most cereal-growing regions and<br />

often become the major cereal leaf disease in wet<br />

and cool conditions. Epidemics are slowed<br />

during periods of dry weather. Serious infections<br />

can cause heavy crop losses if they start early and<br />

affect the uppermost leaves, which contribute<br />

most pho<strong>to</strong>synthetic product <strong>to</strong> the developing<br />

grains. Stagonospora nodorum also infects the<br />

heads of wheat, causing glume blotch. Lesions<br />

develop as small brown spots which enlarge in<strong>to</strong><br />

irregular brownish necrotic areas, giving the leaf<br />

a speckled appearance. Within the necrotic<br />

areas, pycnidia develop beneath the epidermis<br />

which they eventually pierce, releasing a tendril<br />

of cylindrical, three-septate pycnidiospores<br />

(Fig. 17.4d). These can spread the disease <strong>to</strong><br />

neighbouring plants by rainsplash. In S. nodorum,<br />

a second pycnidial state containing minute<br />

unicellular conidia has also been discovered<br />

(Harrower, 1976). These infect host leaves via<br />

germ tubes which penetrate the s<strong>to</strong>mata,<br />

whereas penetration from Stagonospora-type conidia<br />

is directly through the cuticle (Karjalainen &<br />

Lounatmaa, 1986). There is no evidence that<br />

either type of conidium plays a sexual (spermatial)<br />

role. Low temperatures (5 10°C) and irradiation<br />

with ultraviolet light favour conidium<br />

production. Both S. avenae and S. nodorum survive<br />

the winter either on living volunteer plants or as<br />

pycnidia and pseudothecia on stubble. Epidemics<br />

can also be started from infected seeds, and from<br />

ascospores which are released from pseudothecia


PLEOSPORALES<br />

463<br />

Fig17.3 Lep<strong>to</strong>sphaeria acuta.(a)L.S.immature<br />

pseudothecium. Note the young asci (stippled)<br />

elongating between a pre-formed mass of<br />

pseudoparaphyses, and the thin-walled cells<br />

which block the ostiole at this stage but later<br />

dissolve.The centrum subsequently enlarges,<br />

dissolving the pseudoparenchyma surrounding it.<br />

(b) L.S. pycnidium. (c) High-power drawing of cells<br />

lining the pycnidium, showing the origin of the<br />

conidia. (d) Cluster of developing asci from a young<br />

pseudothecium. Note the branching of the<br />

pseudoparaphyses. (e) Stretched bitunicate ascus<br />

showing rupture of the outer wall at its apex.<br />

in the spring and autumn. Good reviews of<br />

S. avenae and S. nodorum have been written by<br />

Eyal (1999) and Cunfer (2000).<br />

An important disease caused by Lep<strong>to</strong>sphaeria<br />

maculans (anamorph Phoma lingam) is blackleg<br />

disease of winter oilseed rape (Brassica napus) and<br />

other brassicas. As the name suggests, the disease<br />

symp<strong>to</strong>ms are seen mainly at the stem bases and<br />

main roots, although the foliage can also be<br />

affected. Infection may occur through leaves,<br />

followed by an endophytic phase lasting several<br />

months, before the blackleg symp<strong>to</strong>ms manifest<br />

themselves. Lesions eventually turn pale and<br />

necrotic, and pycnidia are produced within<br />

them. Where the cortex of infected stems<br />

cracks open, cankers develop. Rain-splashed<br />

pycnidiospores spread the disease during the<br />

growing season. The fungus overwinters on<br />

stubble and infects the new crop by ascospores<br />

released from pseudothecia, although it can also<br />

be seed-borne. The biology of this fungus has been<br />

reviewed by Rouxel and Balesdent (2005).<br />

Lep<strong>to</strong>sphaeria maculans comprises several<br />

morphologically similar species which can be<br />

assigned <strong>to</strong> two groups, the highly destructive<br />

A group which forms cankers and a less damaging<br />

B group which is confined <strong>to</strong> infections of<br />

leaves and the pith (Williams & Fitt, 1999).<br />

The B pathotype has recently been named<br />

L. biglobosa (Shoemaker & Brun, 2001). Some of<br />

the disease symp<strong>to</strong>ms are probably caused by<br />

<strong>to</strong>xins produced by the A pathotype in planta


464 LOCULOASCOMYCETES<br />

Fig17.4 Stagonospora nodorum. (a) Pycnidia seen from above. (b) Pycnidia in section in agar culture. (c) Portion of wall of pycnidium<br />

showing origin of conidia. (d) Conidia. (a,b) <strong>to</strong> same scale; (c,d) <strong>to</strong> same scale.<br />

(Pedras & Biesenthal, 1998; Howlett et al., 2001).<br />

In addition, L. maculans is able <strong>to</strong> de<strong>to</strong>xify phy<strong>to</strong>alexins<br />

produced by the host plant (Pedras et al.,<br />

2000). Since P. lingam is not closely related <strong>to</strong><br />

most other Phoma spp., it has been re-named<br />

Plenodomus (Reddy et al., 1998).<br />

Diverse anamorphic states are known in<br />

other species of Lep<strong>to</strong>sphaeria and Phaeosphaeria,<br />

and some examples of important pathogens are<br />

summarized in Table 17.1.<br />

17.2.3 Ascochyta and Phoma<br />

Several related Ascochyta-type anamorphs infect<br />

legumes, causing diseases such as blight of<br />

chickpea (A. rabiei), foot rot and blight of peas<br />

(A. pinodes), and leaf- and pod-spot of broad beans<br />

(A. fabae) and peas (A. pisi). The pseudothecial<br />

state, where present, is now called Didymella<br />

(formerly Mycosphaerella). These species usually<br />

overwinter on crop residues, infecting new crops<br />

in spring as ascospores, but they are also seedborne.<br />

During the growing season, the infection<br />

is spread by rainsplash of pycnidiospores. An<br />

excellent review of the biology of Ascochyta,<br />

highlighting the severity of diseases which can<br />

be caused, has been written by Pande et al. (2005)<br />

for A. rabiei. Pycnidia contain hyaline, two-celled<br />

conidia which, according <strong>to</strong> Brewer and Boerema<br />

(1965), arise by septation from the conidiogenous<br />

cell. A few conidia with no, two or three<br />

septa may be produced occasionally (Fig. 17.5).<br />

Extensive work has been carried out on characterizing<br />

the mating type idiomorphs of<br />

Ascochyta spp. These data and analyses of other<br />

sequences have shown that Ascochyta spp. belong<br />

<strong>to</strong> the Pleosporales and are not closely related<br />

<strong>to</strong> Mycosphaerella, which is in the Dothideales<br />

(Barve et al., 2003).<br />

Phoma medicaginis (Fig. 17.6) has also been<br />

shown <strong>to</strong> be related <strong>to</strong> Ascochyta spp. (Fatehi et al.,<br />

2003). Unfortunately, little information is available<br />

<strong>to</strong> circumscribe the very large genus Phoma<br />

which is almost certainly polyphyletic, being<br />

associated with several different teleomorphs.<br />

However, a valuable and much-needed monograph<br />

of this difficult group has been written by<br />

Boerema et al. (2004). The conidial fructification<br />

of Phoma is a pycnidium which usually has only<br />

one ostiole, rarely two (Figs. 17.6a and 17.7b).<br />

Like most pycnidia, its outermost wall layer is<br />

pigmented and consists of cells of distinct shape,<br />

which is a useful feature in species identification.<br />

In the case of Phoma spp., it is called a<br />

textura angularis (Fig. 17.6a). The cavity of the<br />

pycnidium is lined by a hymenium of conidiogenous<br />

cells from which one-celled hyaline<br />

pycnidiospores develop. The absence of filiform


PLEOSPORALES<br />

465<br />

Fig17.5 Ascochyta pisi. (a) Pycnidium seen from above, showing a cirrhus of conidia oozing from the ostiole. (b) Conidia<br />

(pycnidiospores). (c) Portion of pycnidium wall in section, showing origin of pycnidiospores.<br />

Fig17.6 Phoma medicaginis. (a) View of the ostiole of a pycnidium from above, showing the angular appearance of the wall surface<br />

comprising a textura angularis. (b) Conidia (pycnidiospores), most of them containing two lipid droplets. (c) Pigmented chlamydospores<br />

formed by old hyphae embedded in agar.The light-refractive globose bodies inside the chlamydospores are lipid droplets.<br />

b-conidia distinguishes Phoma from Phomopsis,<br />

which belongs <strong>to</strong> an al<strong>to</strong>gether different group<br />

(see p. 373). The pycnidiospores often ooze out<br />

from the ostiole as a tendril (cirrhus).<br />

The conidiogenous cells of Phoma are very<br />

small, and details of conidium development are<br />

difficult <strong>to</strong> discern with the light microscope<br />

(Fig. 17.7b). Brewer and Boerema (1965) have<br />

therefore studied spore development with the<br />

electron microscope. They described the process<br />

of spore formation as a monopolar, repetitive<br />

budding of the small, undifferentiated cells of<br />

the pycnidial wall. As repeated spore formation<br />

occurs, the apex of the conidiogenous cell


466 LOCULOASCOMYCETES<br />

Fig17.7 Phoma betae. (a) L.S. pycnidium. (b) Portion of<br />

pycnidium wall, showing conidiogenous cells and conidia.<br />

develops a thickened rim which resembles a<br />

phialide or an annellophore (annellide). Sut<strong>to</strong>n<br />

(1980) has interpreted these spores as<br />

phialoconidia.<br />

Phoma epicoccina is an unusual species because<br />

it possesses a hyphomyce<strong>to</strong>us synanamorph,<br />

Epicoccum nigrum (Fig. 17.8), which is very<br />

common on decaying plant material and in the<br />

soil. The Epicoccum state takes the form of<br />

cushion-shaped sporodochia, which are black or<br />

purplish-red in colour and covered with roughwalled,<br />

warted, segmented, brownish-red conidia<br />

(Fig. 17.8a). The conidia may be violently<br />

projected from the sporodochium, probably by<br />

the rounding off of the two turgid cells on either<br />

side of the septum which separates the conidium<br />

from its conidiophore (Fig. 17.8c; Webster, 1966).<br />

It is possible that conidial discharge is stimulated<br />

by drying because the peak concentration<br />

of Epicoccum spores in air occurs shortly before<br />

noon (Meredith, 1966). The synanamorphic<br />

nature of P. epicoccina and E. nigrum has been<br />

conclusively demonstrated only relatively<br />

recently (Arenal et al., 2000).<br />

17.2.4 Pleospora<br />

Estimates of the number of species of Pleospora<br />

vary, and Kirk et al. (2001) suggested that there<br />

are about 50. Most species form fruit bodies on<br />

moribund herbaceous stems apparently as saprotrophs,<br />

but some are weak pathogens. Of these,<br />

P. bjoerlingii (¼ P. betae) is the cause of blackleg<br />

of sugar beet. Pleospora scirpicola forms its pseudothecia<br />

on the underwater parts of the culms<br />

of the bulrush, Schoenoplectus lacustris, and was<br />

the first fungus in which the ‘jack-in-the-box’<br />

mechanism of discharge of bitunicate asci was<br />

illustrated (see Fig. 17.1).<br />

Pleospora herbarum attacks a wide range of<br />

cultivated hosts, causing such diseases as net<br />

blotch of broad bean and leaf spot of clover,<br />

lucerne and other hosts. It may be seed-borne.


PLEOSPORALES<br />

467<br />

Fig17.8 Epicoccum nigrum. (a) Young sporodochium. (b) Conidiophores and conidia. (c) Conidium almost separated from the<br />

conidiophore. Note the bulging septum of the conidiophore. (d) Two detached conidia.<br />

The pseudothecia are common on overwintered<br />

stems of herbaceous maritime plants. The<br />

large, black, somewhat flattened pseudothecia<br />

contain broad, sac-like bitunicate asci, with<br />

eight yellowish-brown, slipper-shaped ascospores<br />

with transverse and longitudinal septa<br />

(Fig. 17.9a). The anamorphic state, Stemphylium,<br />

produces pigmented muriform conidia and is<br />

often associated with the pseudothecia. The<br />

connection between the teleomorphic and<br />

anamorphic states is readily demonstrated by<br />

shooting ascospores on<strong>to</strong> an agar surface<br />

where they germinate and the resulting mycelium<br />

develops conidia within a few days.<br />

Conversely, cultures started from a single conidium<br />

develop pseudothecia within a few weeks


468 LOCULOASCOMYCETES<br />

Fig17.9 Pleospora herbarum. (a) Ascus and ascospores showing mucilaginous epispore. (b) Stretched bitunicate ascus showing<br />

rupture of outer wall. (c) Developing asci and pseudoparaphyses.The arrows (p) indicate points of branching of ascending and<br />

descending pseudoparaphyses. (d) Conidia of Stemphylium type.<br />

(Weber & Webster, 2000b). The P. herbarum<br />

complex includes a number of similar species<br />

forming conidia which are critically different<br />

from each other in morphology and dimensions<br />

(Simmons, 1969). A distinctive form,<br />

S. vesicarium, is associated with leaf blight of<br />

onions and garlic.<br />

Stemphylium conidia develop singly from the<br />

tips of conidiophores swollen at their apices,<br />

as blown-out ends. A narrow neck of cy<strong>to</strong>plasm<br />

connects the developing spore <strong>to</strong> its conidiophore<br />

through a pore, and Hughes (1953) has<br />

termed conidia of this type porospores<br />

(Fig. 17.9d), but the term poroconidium is also


PLEOSPORALES<br />

469<br />

used (Ellis, 1971b). Electron microscopy studies<br />

(Carroll & Carroll, 1971) have shown that<br />

conidial development is blastic, involving the<br />

whole of the wall at the apex of the conidiogenous<br />

cell. The cy<strong>to</strong>plasmic connection between the<br />

conidiogenous cell and the conidium is narrow,<br />

and is surrounded by two layers of thickened<br />

wall material. Following the detachment of the<br />

first-formed conidium, the conidiophore may<br />

grow out through the detachment scar <strong>to</strong> form<br />

a second conidium, a process described as<br />

percurrent conidiogenesis. The conidia of<br />

P. herbarum are formed more readily in cultures<br />

illuminated by near-UV light (Leach, 1968),<br />

whereas daylight and low temperature stimulate<br />

pseudothecial development (Leach, 1971). The<br />

fungus is homothallic. According <strong>to</strong> Meredith<br />

(1965) the conidia are violently jolted from the<br />

tip of the conidiophores.<br />

17.2.5 Lewia<br />

The genus Lewia was named in honour of L. E.<br />

Wehmeyer by Simmons (1986) for Pleospora-like<br />

fungi with Alternaria anamorphs. Six ascocarpic<br />

species have been recognized, fruiting on grasses<br />

(including cereals) and on dicotyledonous hosts<br />

(including Brassica and Pastinaca) (Kwasna &<br />

Kosiak, 2003). The separation of Lewia from<br />

Pleospora is supported by molecular evidence<br />

(Pryor & Gilbertson, 2000).<br />

Lewia infec<strong>to</strong>ria (¼ Pleospora infec<strong>to</strong>ria) forms<br />

black, shining, subepidermal pseudothecia on<br />

overwintered grass and cereal culms. It has<br />

golden-brown muriform ascospores with up <strong>to</strong><br />

five transverse septa. The central cells of the<br />

ascospores also contain one or rarely two longitudinal<br />

septa (Fig. 17.10a). In culture, this<br />

fungus forms branching chains of obclavate,<br />

brown-coloured (melanized), muriform, beaked<br />

spores (dictyospores or dictyoconidia) and new<br />

spores are formed at the tip of the chain<br />

(Fig. 17.10c). A darkly pigmented thickened<br />

annulus is visible at the base of the conidium<br />

and at the apex of the conidiophore surrounding<br />

the point of spore separation and, if the spore<br />

has occupied an intercalary position on the spore<br />

chain, there is also an annulus at the opposite<br />

end. Chain branching occurs where a conidium<br />

produces more than one spore. Conidia of this<br />

type have been classified in the anamorph genus<br />

Alternaria, and are poroconidia. The conidial state<br />

of L. infec<strong>to</strong>ria is A. infec<strong>to</strong>ria.<br />

17.2.6 Alternaria<br />

About 50 species of Alternaria are known which<br />

have not been connected <strong>to</strong> a teleomorph<br />

(Neergaard, 1945; Joly, 1964; Simmons, 1986;<br />

Kirk et al., 2001). The taxonomy of Alternaria is<br />

difficult. Simmons (1992) has given a key <strong>to</strong> 10<br />

species-groups and Ellis (1971a, 1976) has<br />

described and figured some common species.<br />

Despite the absence of formal evidence for sexual<br />

reproduction in many species of Alternaria,<br />

Berbee et al. (2003) have shown that three species<br />

of Alternaria not known <strong>to</strong> have sexual states,<br />

A. brassicae, A. brassicicola and A. tenuissima, have<br />

mating type gene sequences. In any one isolate of<br />

these species, only one mating type idiomorph<br />

was found, but in other isolates of the same<br />

species the opposite idiomorph was detected.<br />

This suggests that these currently asexual species<br />

were derived from sexually reproducing<br />

ances<strong>to</strong>rs.<br />

The fine structure of conidial development<br />

from a pre-existing conidium in A. brassicae has<br />

been studied by Campbell (1968). The mature<br />

spore has a two-layered wall, the outer of which<br />

is melanized. A pore develops in the outer wall,<br />

probably by enzymatic activity, and the inner<br />

wall layer expands through the pore <strong>to</strong> become<br />

the primary wall of the new conidium. Later, this<br />

in turn becomes two-layered. Transverse and<br />

longitudinal septa develop within the spore,<br />

but these are incomplete; a pore in each<br />

septum allows cy<strong>to</strong>plasmic continuity between<br />

adjacent cells and flow of cy<strong>to</strong>plasm through the<br />

spore <strong>to</strong> provide material for the formation of<br />

new spores at the tip of the chain.<br />

The shape of the conidium in Alternaria<br />

affects its aerodynamic properties. Several<br />

species have conidia with long beaks, e.g.<br />

A. solani and A. brassicae (Fig. 17.11). It has been<br />

suggested that the long beaks increase the<br />

chance of wind-dispersal as compared <strong>to</strong> species<br />

with smaller, non-beaked conidia (Chou & Wu,<br />

2002). Long beaks also increase the drag on the


470 LOCULOASCOMYCETES<br />

Fig17.10 Lewiainfec<strong>to</strong>ria. (a) Asci, one intact, the<br />

other stretched by expansion of the endotunica.<br />

(b) Germinating ascospore. (c) Alternaria<br />

infec<strong>to</strong>ria conidial state developed in culture<br />

from a germinating ascospore. (d) A conidium<br />

from an intercalary position in a spore chain<br />

showing a scar at each end of the conidium.<br />

(e) A conidium from the end of a spore chain<br />

with only one basal scar.<br />

spore, reducing its settling velocity (McCartney<br />

et al., 1993).<br />

Some species of Alternaria are of considerable<br />

economic significance (Chelkowski & Visconti,<br />

1992; Rotem, 1994). Alternaria alternata (¼ A.<br />

tenuis) is the name given <strong>to</strong> a widespread and<br />

cosmopolitan opportunistic saprotroph, reported<br />

on all kinds of senescent plant material.<br />

Unfortunately, because of difficulties in identification,<br />

it is probable that the name encompasses<br />

several distinct taxa (Roberts et al., 2000).<br />

Alternaria spp. are associated with diseases of<br />

crops, often showing a degree of host specificity<br />

as indicated in Table 17.2. Most of these diseases<br />

are seed-borne; seeds become infected from the<br />

flowering stage onwards (Rotem, 1994). The hostspecific<br />

pathogens may have evolved from nonspecific<br />

saprotrophic forms of ‘A. alternata’ by the<br />

selection and multiplication of mutants capable<br />

of secreting host-specific <strong>to</strong>xins (Scheffer, 1992;<br />

Rotem, 1994; Thomma, 2003). Many of these<br />

host-specific <strong>to</strong>xins have been characterized<br />

chemically (Otani & Kohmo<strong>to</strong>, 1992). Numerous<br />

myco<strong>to</strong>xins are also produced by Alternaria spp.<br />

(Montemurro & Visconti, 1992), and some have<br />

severe and fatal consequences if they accumulate


PLEOSPORALES<br />

471<br />

Fig17.11<br />

Alternaria brassicae. (a) Conidiophores and conidia. (b) Conidia. Note the long conidial beaks.<br />

in human food and animal feed. Spores of<br />

Alternaria spp. are abundant in the air in late<br />

summer and autumn and may be a cause of<br />

inhalant allergy causing asthma. Some species,<br />

e.g. ‘A. alternata’, A. infec<strong>to</strong>ria and A. longipes are<br />

rare opportunistic human pathogens, associated<br />

with diseases of bone, cutaneous tissue, ears,<br />

eyes, nose and the urinary tract (Schell, 2003).<br />

Other species are insect pathogens.<br />

17.2.7 Cochliobolus<br />

In Cochliobolus, the pseudothecium is long-necked<br />

and contains elongate, transversely septate<br />

ascospores spirally coiled around each other<br />

within a vestigially bitunicate ascus (see<br />

Figs. 8.15b d).<br />

Anamorphs associated with Cochliobolus<br />

The genus Cochliobolus contains some of the<br />

best-studied and most highly damaging plant<br />

pathogens. As in other Loculoascomycetes,<br />

the taxonomic his<strong>to</strong>ry of this genus has<br />

been <strong>to</strong>rtuous (Sivanesan, 1987; Alcorn, 1988);<br />

the anamorphs were formerly classified in<br />

Helminthosporium, then transferred <strong>to</strong> Drechslera,<br />

and are now called Bipolaris (Fig. 17.13) and


472 LOCULOASCOMYCETES<br />

Table 17.2. Selected plant-pathogenic species of Alternaria (from Holliday, 1998).<br />

Species of Alternaria Disease Comments<br />

‘A. alternata’<br />

Widerangeofhostsanddiseases<br />

‘A. alternata’f. sp. lycopersici Toma<strong>to</strong> stem canker<br />

A. brassicae Brassica greyleaf spot Seed-borne<br />

A. brassicicola Brassica blackleaf spot Seed-borne<br />

A. carthami Safflower leaf spot Seed-borne<br />

A. dauci Carrot leaf blight Seed-borne<br />

A. dianthi Carnationleaf blight Seed-borne<br />

A. linicola Linseed seedling blight Seed-borne<br />

A. macrospora Cot<strong>to</strong>nleaf spot Seed-borne<br />

A. mali Apple core rot<br />

A. porri Onion purple blotch Seed-borne<br />

A. radicina Carrot black rot; also infects celery and parsnip<br />

A. solani Pota<strong>to</strong> early blight and tuber rot<br />

Toma<strong>to</strong> early blight and fruit rot<br />

Curvularia (Fig. 17.14). The revised genus<br />

Helminthosporium is now rather small, with<br />

about 20 species and H. velutinum as the typespecies<br />

(Fig. 17.12). It has affinities <strong>to</strong><br />

Lep<strong>to</strong>sphaeria (Olivier et al., 2000). Drechslera, in<br />

contrast, is the anamorphic state of Pyrenophora<br />

(Fig. 17.16; see p. 477). All these anamorphic<br />

forms produce pigmented (melanized) spores<br />

with only transverse septa, but they differ in<br />

their pattern of conidiogenesis and in the<br />

ultrastructural appearance of their cell walls.<br />

Sporogenesis by Helminthosporium is tretic, i.e.<br />

porospores are formed as described for<br />

Stemphylium by the digestion of a small hole<br />

in<strong>to</strong> the conidiophore cell wall, followed by an<br />

extension of the inner wall <strong>to</strong> form the conidium<br />

which then lays down its outer wall (Fig. 17.12c).<br />

In contrast <strong>to</strong> Pleospora, each site can produce<br />

only one conidium, and conidia can develop<br />

apically or laterally on the conidiophore. If the<br />

conidium is produced apically, the conidiophore<br />

cannot grow further. The sequence of conidium<br />

development in Bipolaris and Curvularia is not<br />

of this type. In Bipolaris, the first conidium<br />

always develops apically, and subsequent<br />

conidia are formed either by growing through<br />

the scar left by the first conidium, or by the<br />

conidiophore growing past the first conidium <strong>to</strong><br />

form a new apex producing the second conidium<br />

(B. sorokiniana; Fig. 17.13b). Bipolaris is so named<br />

because the conidia germinate by emitting two<br />

germ tubes, one at either end (Fig. 17.13d), and<br />

these grow as extensions of the long axis of the<br />

spore. This is in contrast <strong>to</strong> Helminthosporium<br />

conidia in which each cell is principally capable<br />

of germination, and the germ tubes grow<br />

perpendicular <strong>to</strong> the long axis of the spore.<br />

In Curvularia, the conidia are curved because<br />

of an unevenly swollen central cell. The end cells<br />

are usually less strongly pigmented than the<br />

central cell (Figs. 17.14b,c). The first conidium<br />

also develops at the apex of the conidiophore as<br />

a poroconidium extending through a tiny pore,<br />

and then the conidiophore develops a new<br />

subterminal growing point from which a<br />

second conidium initial arises. The process is<br />

repeated so that a succession of new apices, each<br />

terminated by a conidium, is formed (Fig. 17.14a).<br />

The term sympodula has been applied <strong>to</strong> such<br />

a conidiophore producing conidia sympodially<br />

(Kendrick & Cole, 1968). In some Curvularia spp.,<br />

such as C. cymbopogonis (Fig. 17.14c), the base of<br />

the conidium bears a protuberant hilum.<br />

A feature of the conidia of Drechslera,<br />

Helminthosporium and Bipolaris, but not<br />

Curvularia, is that they are dis<strong>to</strong>septate<br />

(Luttrell, 1963). This means that the wall separating<br />

adjacent conidial segments is visibly


PLEOSPORALES<br />

473<br />

Fig17.12 Helminthosporium velutinum. (a) Conidiophores and conidia. (b) Detached conidia. (c) Details of conidial development.<br />

Note the narrow channels in the wall through which cy<strong>to</strong>plasm passes <strong>to</strong> the developing conidia.This type of development is tretic.<br />

(b)and(c)<strong>to</strong>samescale.<br />

different from the outer wall surrounding the<br />

entire conidium (White et al., 1973). This can be<br />

readily seen even with the light microscope<br />

because the individual conidial cells of<br />

Drechslera, Helminthosporium and Bipolaris appear<br />

like peas in a pod (Figs. 17.12, 17.13, 17.16). In<br />

Curvularia, the conidial septa are of the normal<br />

(euseptate) type (Fig. 17.14).


474 LOCULOASCOMYCETES<br />

Fig17.13 Bipolaris sorokiniana, the conidial state of Cochliobolus sativus. (a) Developing conidia.The arrows point <strong>to</strong> developing septa.<br />

(b) Conidiophore showing the development of a second conidium lateral <strong>to</strong> the first. (c) Mature conidia. (d) Two germinating<br />

conidia showing emergence of germ tubes from each end of the conidium.


PLEOSPORALES<br />

475<br />

Fig17.14 Curvularia spp. (a) Curvularia lunata, the conidial state of Cochliobolus lunatus.Conidiophores showing sequence of conidial<br />

development. (b) Mature detached conidia of C. lunata. Note the paler end cells. (c) Mature detached conidia of C. cymbopogonis<br />

showing the protuberant hilum.<br />

The pathology of Cochliobolus<br />

Cochliobolus has been examined by DNA sequencing<br />

methods (Berbee et al., 1999), which revealed<br />

two separate groups. All important pathogens<br />

belong <strong>to</strong> one group and have Bipolaris<br />

anamorphs. In general, the teleomorphs are<br />

rare or absent in nature, and the diseases are<br />

carried mainly by the thick-walled conidia,<br />

which can survive in the soil but can also infect<br />

seeds. Cochliobolus sativus (B. sorokiniana) is the<br />

cause of a variety of root and leaf necroses of<br />

cereals (especially wheat and barley) in warm<br />

humid climates of South East Asia, Australia,<br />

North and South America. A good review of its<br />

biology has been written by Kumar et al. (2002).<br />

It is a typical hemibiotrophic pathogen which<br />

forms quiescent infections in the first-infected<br />

epidermal cell, followed by a necrotrophic phase<br />

in which the surrounding tissue is aggressively<br />

invaded. This pathogen shares with Botrytis<br />

cinerea (see p. 435) the ability <strong>to</strong> evade the<br />

oxidative burst launched as a defence response<br />

by the newly infected host (Kumar et al., 2001).<br />

Several sesquiterpene-type <strong>to</strong>xins are produced<br />

and contribute <strong>to</strong> a weakening of the host;<br />

prehelminthosporol (Fig. 17.15a) seems <strong>to</strong> be<br />

the most potent. These <strong>to</strong>xins act in a nonspecific<br />

manner on several different cellular<br />

processes. Interestingly, barley cultivars possessing<br />

the mlo-type resistance against powdery


476 LOCULOASCOMYCETES<br />

Fig17.15 Toxins produced by Cochliobolus spp. a.The sesquiterpene prehelminthosporol produced by C. sativus.b.Thechlorinated<br />

pentapeptide vic<strong>to</strong>rin C produced by C. vic<strong>to</strong>riae. c.The dominant form of T-<strong>to</strong>xin, a polyketide produced by C. heterostrophus.<br />

d.The major form of HC-<strong>to</strong>xin, a tetrapeptide produced by C. carbonum.<br />

Fig17.16 Pyrenophora tritici-repentis. (a) Immature pseudothecial stroma on overwintered barley stubble, with conidia produced<br />

on the setae and on separate conidiophores. (b,c) Conidia of Drechslera tritici-repentis mounted in lactic acid (b) showing the typical<br />

‘peas-in-a-pod’ appearance, and mounted in water (c). (b) and (c) <strong>to</strong> same scale.<br />

mildew (see p. 408) are highly susceptible<br />

<strong>to</strong> C. sativus. A hypersensitive response which<br />

effectively suppresses biotrophic pathogens may<br />

thus actually enhance pathogenicity by others<br />

with a necrotrophic potential.<br />

A similar feature was observed with<br />

C. vic<strong>to</strong>riae, which became pathogenic on the<br />

Vic<strong>to</strong>ria cultivar of oats bred for resistance<br />

against the crown rust, Puccinia coronata<br />

(Meehan & Murphy, 1946). The main <strong>to</strong>xin<br />

produced by C. vic<strong>to</strong>riae, vic<strong>to</strong>rin C (Fig. 17.15b),<br />

is one of several related cyclic pentapeptides<br />

which are absolutely essential and sufficient for<br />

pathogenesis, i.e. the purified <strong>to</strong>xin can reproduce<br />

the disease symp<strong>to</strong>ms which consist of<br />

widespread leaf necrosis, and strains of<br />

C. vic<strong>to</strong>riae not synthesizing this <strong>to</strong>xin do not<br />

cause the disease. It now seems that vic<strong>to</strong>rin<br />

triggers a form of hypersensitive response, i.e.<br />

widespread host cell death (Navarre & Wolpert,


PLEOSPORALES<br />

477<br />

1999). Vic<strong>to</strong>rin binds <strong>to</strong> mi<strong>to</strong>chondrial proteins,<br />

and it is possible that it initiates cell death via<br />

mi<strong>to</strong>chondrial dysfunction (Curtis & Wolpert,<br />

2002). Intriguingly, in the mammalian equivalent<br />

of the hypersensitive response (i.e. apop<strong>to</strong>sis),<br />

mi<strong>to</strong>chondria are also among the first<br />

organelles <strong>to</strong> break down.<br />

Mi<strong>to</strong>chondria are also the target of the<br />

specific <strong>to</strong>xin produced by C. heterostrophus<br />

(B. maydis), which caused the well-described<br />

southern leaf blight of corn in the United<br />

States in 1970 (Ullstrup, 1972; Schuman, 1991).<br />

Prior <strong>to</strong> that epidemic, maize breeders had relied<br />

heavily on male-sterile cultivars, i.e. cultivars<br />

which do not produce viable pollen and are<br />

therefore dependent on pollen from another<br />

cultivar for seed production. This facilitated<br />

the production of hybrids, i.e. the F1 progeny of<br />

genetically dissimilar parents. These often<br />

produce particularly high yields, a phenomenon<br />

known as hybrid vigour. Male sterility was<br />

achieved by a mi<strong>to</strong>chondrial mutation called<br />

cms-T (cy<strong>to</strong>plasmic male sterility). The plant line<br />

used as the pollen donor did not contain the<br />

cms-T mutation, so that the hybrid seeds grew<br />

in<strong>to</strong> plants capable of producing pollen in the<br />

field of the farmers. In 1970, about 80% of the<br />

maize crop contained cms-T cy<strong>to</strong>plasm. At about<br />

the same time a mutation in a minor leaf<br />

spot pathogen, C. heterostrophus race O, spread<br />

in the field. This new race produced several<br />

related polyketides collectively called T-<strong>to</strong>xin<br />

(Fig. 17.15c), which specifically affected the<br />

mi<strong>to</strong>chondria of hybrid maize. The epidemic<br />

of 1970 resulted in crop losses <strong>to</strong>talling over<br />

US $1 billion for that year. The target site of<br />

T-<strong>to</strong>xin in cms-T maize mi<strong>to</strong>chondria is now<br />

known <strong>to</strong> reside in a small protein present<br />

as a tetramer in the outer mi<strong>to</strong>chondrial membrane.<br />

T-<strong>to</strong>xin binds directly <strong>to</strong> this protein,<br />

causing conformational changes which open<br />

up pores and render mi<strong>to</strong>chondria leaky<br />

(Levings et al., 1995; Wolpert et al., 2002). Race O<br />

causes only minor leaf spots on cms-T as well as<br />

other cultivars of maize.<br />

Cochliobolus carbonum is the cause of northern<br />

leaf spot and ear rot of maize. There are three<br />

races, of which only race 1 is a serious pathogen<br />

on all those cultivars of maize containing two<br />

recessive alleles of the Hm1 resistance gene. This<br />

strongly enhanced pathogenicity is caused by a<br />

group of cyclic tetrapeptides collectively called<br />

HC-<strong>to</strong>xin (Fig. 17.15d), where HC stands for<br />

Helminthosporium carbonum, the former name of<br />

the anamorph. Resistant cultivars produce an<br />

enzyme which de<strong>to</strong>xifies HC-<strong>to</strong>xin. The susceptibility<br />

of certain maize cultivars was caused<br />

by a simultaneous inactivation of both duplicate<br />

genes encoding the enzyme, HC-<strong>to</strong>xin reductase<br />

(Multani et al., 1998). The mode of action of HC<strong>to</strong>xin<br />

is not yet clear; it seems <strong>to</strong> inhibit, rather<br />

than induce, defence responses (Wolpert et al.,<br />

2002).<br />

As we have seen, certain strains of<br />

C. carbonum, C. heterostrophus and C. vic<strong>to</strong>riae have<br />

caused catastrophic epidemics on particular<br />

cereal hosts employing biochemically unrelated<br />

<strong>to</strong>xins. This specificity of action i.e. a specific<br />

pathogen race being pathogenic only against<br />

certain host cultivars paved the way <strong>to</strong>wards<br />

an understanding of the gene-for-gene concept<br />

(see pp. 112 and 397). A further question of<br />

interest is the origin of these highly aggressive<br />

strains. A partial answer was found somewhat<br />

fortui<strong>to</strong>usly by an examination of the mating<br />

type genes in all three pathogens. Both idiomorphs<br />

MAT-1 and MAT-2 have been found in<br />

various field isolates of C. heterostrophus and<br />

C. carbonum, but all known isolates of C. vic<strong>to</strong>riae<br />

belong <strong>to</strong> MAT-2. Further, C. vic<strong>to</strong>riae and MAT-1<br />

strains of C. carbonum are interfertile. These<br />

observations have led <strong>to</strong> the suggestion that<br />

C. vic<strong>to</strong>riae arose from a MAT-2 strain of C.<br />

carbonum which received the gene cluster for<br />

pathogenicity on oats (i.e. the genes encoding the<br />

enzymes necessary for <strong>to</strong>xin synthesis) by horizontal<br />

gene transfer (Christiansen et al., 1998).<br />

The integration of this gene cluster must have<br />

been close <strong>to</strong> the MAT-2 locus, so that it did<br />

not spread <strong>to</strong> MAT-1 strains of C. carbonum by<br />

crossing-over.<br />

17.2.8 Pyrenophora (anamorph Drechslera)<br />

In its taxonomically restricted use, the Drechslera<br />

state (Fig. 17.16b) is the conidial form of<br />

Pyrenophora and is clearly defined as a monophyletic<br />

group (Zhang & Berbee, 2001). Drechslera


478 LOCULOASCOMYCETES<br />

is commonly found in the field, whereas<br />

Pyrenophora-type pseudothecia are uncommon.<br />

Species belonging <strong>to</strong> this group are pathogens<br />

of cereals and grasses, and some of them cause<br />

significant diseases in agricultural situations.<br />

The disease symp<strong>to</strong>ms are similar <strong>to</strong> those<br />

caused by other members of the Pleosporales,<br />

and phy<strong>to</strong><strong>to</strong>xic substances are produced by<br />

several members of the genus.<br />

Pyrenophora tritici-repentis (anamorph Drechslera<br />

tritici-repentis) occurs on a range of grasses,<br />

including, as the name suggests, Agropyron<br />

(formerly Triticum) repens and wheat (Triticum<br />

aestivum). The disease caused is known as yellow<br />

leaf spot or tan spot of wheat (De Wolf et al.,<br />

1998). Pyrenophora tritici-repentis is spread as seedborne<br />

infections but also overwinters on infected<br />

stubble, which is the most important source of<br />

inoculum, giving rise <strong>to</strong> pseudothecia and<br />

conidia in spring. Pseudothecia can be identified<br />

by their large size and by the presence of dark<br />

setae around the pseudothecial neck (Fig. 17.16a).<br />

The ascospores are transversely septate, with a<br />

longitudinal septum also present in one of the<br />

central cells. The conidia of Drechslera triticirepentis<br />

are very large and have a variable<br />

number of dis<strong>to</strong>septa (Fig. 17.16b). They are<br />

sometimes produced on the pseudothecial<br />

setae, or they arise directly from stubble or<br />

from necrotic leaf lesions. Phy<strong>to</strong><strong>to</strong>xins are<br />

involved in causing disease symp<strong>to</strong>ms. Most<br />

unusually, they consist of at least two extracellular<br />

proteins synthesized by ribosomes<br />

(Wolpert et al., 2002). They are a critical fac<strong>to</strong>r<br />

in determining the host specificity of infections.<br />

Pyrenophora teres is common wherever barley<br />

is grown, and is the major barley pathogen,<br />

especially in humid regions. This species exists in<br />

two forms which are distinguished by their<br />

symp<strong>to</strong>ms, P. teres f. teres causing net blotch on<br />

barley leaves and P. teres f. maculata causing<br />

brown leaf spots. These two forms can hybridize<br />

in the labora<strong>to</strong>ry and also in the field (Campbell<br />

& Crous, 2003). In addition <strong>to</strong> ascospores and<br />

Drechslera-type macroconidia similar <strong>to</strong> those of<br />

P. tritici-repentis, a pycnidial state producing<br />

unicellular conidia is also apparently associated<br />

with P. teres, although its role in the disease<br />

cycle is uncertain (Smith et al., 1988). The<br />

epidemiology of the disease is similar <strong>to</strong> P.<br />

tritici-repentis, as is the involvement of phy<strong>to</strong><strong>to</strong>xins<br />

in causing leaf necrosis. Leaf chlorosis and<br />

necrosis can be reproduced by phy<strong>to</strong><strong>to</strong>xins<br />

purified from cultures of P. teres. However,<br />

biochemically the phy<strong>to</strong><strong>to</strong>xins involved are<br />

rather different, the most potent of them being<br />

the aspartic acid derivative aspergillomarasmine<br />

A (Weiergang et al., 2002).<br />

17.2.9 Venturia<br />

The genus Venturia contains some 50 species<br />

which cause scabs, i.e. limited lesions with a<br />

scurfy appearance, on the leaves and fruits of<br />

various trees. The genus is an unusual member<br />

of the Pleosporales in producing asci with oneseptate<br />

ascospores, but Silva-Hanlin and Hanlin<br />

(1999) have confirmed its position within this<br />

order. The most important species is V. inaequalis<br />

which parasitizes apple (Malus spp.) and hosts<br />

related <strong>to</strong> it. This fungus is cosmopolitan and<br />

extremely common on apple fruits and leaves if<br />

fungicide treatments are not carried out<br />

(Fig. 17.17a). In many regions, scab is the most<br />

serious apple disease. The fungus overwinters on<br />

fallen leaves which, in spring, give rise <strong>to</strong><br />

pseudothecia (Fig. 17.17b) releasing ascospores<br />

during periods of wetness. The ascospores<br />

require surface wetness in order <strong>to</strong> infect apple<br />

leaves. Infection is mediated by appressoria, but<br />

the invasion is limited <strong>to</strong> the space between the<br />

cuticle and the epidermis; the latter is not<br />

pierced, and haus<strong>to</strong>ria are not formed. In this<br />

way, the fungus persists for several weeks.<br />

Conidia are eventually produced from such<br />

subcuticular stromata, and they spread the<br />

disease during the growing season. Invasion of<br />

host tissue takes place only on dead leaves in the<br />

autumn when V. inaequalis switches <strong>to</strong> a saprotrophic<br />

growth phase and produces pseudothecial<br />

initials. The biology of Venturia, which<br />

has been summarized by MacHardy et al. (2001),<br />

is thus very unusual for members of the<br />

Pleosporales.<br />

From the above summary of the infection<br />

biology of Venturia it is apparent that the key <strong>to</strong><br />

apple scab management lies in controlling the<br />

ascospore inoculum in spring. One commonly


PLEOSPORALES<br />

479<br />

Fig17.17 Venturia inaequalis. (a) Apples showing scab symp<strong>to</strong>ms.<br />

(b) Section through a pseudothecium in an overwintered<br />

apple leaf.<br />

practised approach is disease forecasting,<br />

based on the knowledge that ascospore<br />

discharge occurs within 1 2 h of wetting ripe<br />

pseudothecia, and that infection requires leaf<br />

surface wetness for some 25 h at 6°C or9hat<br />

16 24°C (Smith et al., 1988). Under certain<br />

circumstances, e.g. after a prolonged dry<br />

period, some time will elapse before the pseudothecia<br />

have produced a fresh crop of ascospores<br />

after the onset of rain, and this can be<br />

integrated in<strong>to</strong> forecasting systems (Stensvand<br />

et al., 2005). Protective fungicide sprays have <strong>to</strong> be<br />

applied as soon as possible after the onset of<br />

conditions conducive <strong>to</strong> infection, and especially<br />

if a high density of air-borne ascospores has<br />

already been detected by spore traps or other<br />

means (Kollar, 1998). Curative fungicides can be<br />

applied one or a few days after infection.<br />

Numerous fungicides are in use against apple<br />

scab. Protective agents include copper-based<br />

formulations which are registered in some<br />

countries even for organic farming, or the thiol<br />

reactant captan. Important curative fungicides<br />

include strobilurins (respiration inhibi<strong>to</strong>rs), and<br />

myclobutanil and imidazoles (demethylation<br />

inhibi<strong>to</strong>rs of ergosterol biosynthesis).<br />

A different control strategy is <strong>to</strong> reduce the<br />

available ascospore inoculum in the spring by<br />

encouraging the decomposition of leaves during<br />

winter. This can be achieved by applying urea <strong>to</strong><br />

the leaf litter, or by using a flail mower <strong>to</strong> shred<br />

the leaves (Sut<strong>to</strong>n et al., 2000). It may also prove<br />

possible <strong>to</strong> spray leaves before leaf-fall with<br />

spores of fungi antagonistic <strong>to</strong> Venturia (Carisse<br />

et al., 2000). Yet another approach is the breeding<br />

of resistant cultivars.<br />

17.2.10 Sporormiella<br />

There are about 70 species of Sporormiella (Ahmed<br />

& Cain, 1972). Molecular studies indicate that the<br />

genus has affinities with Pleosporales (Liu et al.,<br />

1999). Most species form pseudothecia on the<br />

dung of herbivores, but some are isolated from<br />

the soil or as endophytes. Characteristic features<br />

are dark, transversely septate ascospores which<br />

may disarticulate in<strong>to</strong> separate part-spores, each<br />

of which is capable of germination, and whose<br />

walls are often marked by a hyaline longitudinal<br />

or oblique germ slit. Sporormiella intermedia is one<br />

of the common species and has thin transparent<br />

pseudothecial walls through which asci can be<br />

seen (Fig. 17.18a). The ascospores of S. intermedia<br />

are four-celled and surrounded by a mucilaginous<br />

envelope. Spore discharge is nocturnal<br />

(Walkey & Harvey, 1966b).<br />

Because of the unmistakable shape of its<br />

ascospores and its association with dung,<br />

Sporormiella has been used in an archaeological


480 LOCULOASCOMYCETES<br />

Fig17.18 Sporormiella intermedia. (a) Pseudothecium with asci visible through the transparent wall. (b) Ripe unextended ascus<br />

showing the double wall. (c) Elongating ascus showing rupture of the outer wall (ec<strong>to</strong>tunica) and extension of the inner (endotunica).<br />

(d) Ascospore separated in<strong>to</strong> its four component cells. (e) Intact ascospore.<br />

context as an indica<strong>to</strong>r of changes in vegetation<br />

and land management. Thus, Burney et al. (2003)<br />

have shown for the island of Madagascar that<br />

Sporormiella was very common before the settlement<br />

of humans which occurred after ad 200<br />

and then declined in abundance along with the<br />

extinction of several groups of large herbivores.<br />

After ad 1100, spore densities of Sporormiella<br />

in sediments showed an increase, coinciding<br />

with the introduction of grazing domesticated<br />

lives<strong>to</strong>ck.<br />

17.3 Dothideales<br />

This is a large group of ascomycetes containing<br />

an enormous variety of conidial forms, both<br />

hyphomyce<strong>to</strong>us and pycnidial. Where present,<br />

the teleomorph consists of dark-celled pseudothecial<br />

ascomata, usually developing as<br />

locules within an ascostroma. The asci are<br />

bitunicate (fissitunicate). In contrast <strong>to</strong> the<br />

Pleosporales, inter-ascal tissue (hamathecium<br />

with pseudoparaphyses) is generally lacking. A<br />

relationship between the Dothideales and the<br />

Pleosporales has been suggested on the basis of<br />

molecular data (Silva-Hanlin & Hanlin, 1999;<br />

Lumbsch & Lindemuth, 2001). Many members<br />

are saprotrophic on dead plant material, but<br />

some grow as endophytes and some are plant<br />

pathogens. The single ascocarpic example which<br />

we shall study is Mycosphaerella.<br />

17.3.1 Mycosphaerella<br />

Mycosphaerella is one of the largest genera of<br />

ascomycetes, containing over 2000 described<br />

species (Corlett, 1991). However, many of the<br />

names are based mainly on the association of a<br />

Mycosphaerella with a particular host plant. Given<br />

the lack of critical inoculation experiments <strong>to</strong><br />

clarify their host range, mating experiments or<br />

DNA sequence comparisons with similar forms<br />

on other plants, it is likely that many names<br />

are synonyms. Some species are plurivorous,<br />

growing on a broad range of monocotyledonous<br />

and dicotyledonous hosts. Many species of<br />

Mycosphaerella cause diseases of economic significance,<br />

and some of them are listed in Table 17.3.<br />

Most of these diseases involve the necrosis of<br />

host plant tissue, and the <strong>to</strong>xins produced by<br />

the pathogens are commonly associated with<br />

the disease symp<strong>to</strong>ms (e.g. Cercospora beticola; see<br />

p. 481).


DOTHIDEALES<br />

481<br />

Table 17.3. Some anamorph genera with Mycosphaerella teleomorphs.<br />

Mycosphaerella Anamorph Diseases caused<br />

M. graminicola Sep<strong>to</strong>ria tritici Leaf blotch of wheat<br />

M. brassicicola Asteromella brassicae Ring spot of brassicas<br />

M. tassiana Cladosporium herbarum<br />

M. berkeleyi Passalora personata Groundnut defoliation<br />

(unknown) Cercospora beticola Leaf spo<strong>to</strong>f sugar beet<br />

M. fijiensis Paracercospora fijiensis Leaf spot of banana<br />

M. musicola Pseudocercospora musae Siga<strong>to</strong>ka disease of banana<br />

Crous et al. (2000, 2001) and Goodwin and<br />

Zismann (2001) listed and discussed the bewildering<br />

diversity of anamorph genera (about 23)<br />

connected with Mycosphaerella teleomorphs, and<br />

performed phylogenetic analyses on representatives<br />

of most of them. The list includes pycnidial<br />

forms such as Sep<strong>to</strong>ria and Asteromella, but also<br />

numerous hyphomyce<strong>to</strong>us form-genera such as<br />

Cercospora, Pseudocercospora and Cladosporium<br />

(Table 17.3). Despite this range of anamorphs,<br />

molecular evidence has somewhat surprisingly<br />

indicated that the genus Mycosphaerella is monophyletic<br />

(Crous et al., 2000, 2001; Goodwin &<br />

Zismann, 2001). Pseudothecia of Mycosphaerella<br />

are globose and small, rarely more than 100 mm<br />

in diameter. Because they show relatively little<br />

variation, they are difficult <strong>to</strong> identify <strong>to</strong> species<br />

level. Pseudothecia develop subepidermally,<br />

usually on leaves. The asci develop in a basal<br />

fascicle. The ascospores are hyaline with a single<br />

transverse septum (Fig. 17.19). In these features<br />

Mycosphaerella is similar <strong>to</strong> Venturia, although<br />

these two genera are not closely related.<br />

17.3.2 Mycosphaerella graminicola<br />

(anamorph Sep<strong>to</strong>ria tritici)<br />

The leaf blotch disease of wheat caused by<br />

M. graminicola (Fig. 17.20) is very similar <strong>to</strong> that<br />

caused by the wheat strain of Phaeosphaeria<br />

nodorum (see p. 17.2.2), and the two diseases<br />

often co-occur on wheat crops and are controlled<br />

in the same way, especially by the application of<br />

fungicides. The most important fungicides are<br />

strobilurin-type compounds and ergosterol<br />

biosynthesis inhibi<strong>to</strong>rs. The ascospores of<br />

M. graminicola, produced from pseudothecia<br />

initially on overwintering stubble and later<br />

from infected leaves, are the main source of<br />

inoculum, and conidia are thought <strong>to</strong> be of lesser<br />

importance as propagules of the disease (Eyal,<br />

1999). In consequence, the genetic diversity of<br />

M. graminicola in the field is often very high,<br />

with one square metre of infected wheat shown<br />

<strong>to</strong> contain about 70 genetically different strains<br />

(Zhan et al., 2001). Infection by germinating<br />

ascospores and conidia of M. graminicola is<br />

almost always through s<strong>to</strong>mata (Duncan &<br />

Howard, 2000), in contrast <strong>to</strong> Phaeosphaeria<br />

nodorum where it occurs directly through the<br />

cuticle. Following penetration, intercellular<br />

colonization of the surrounding leaf tissue by<br />

hyphae of M. graminicola occurs, but the onset<br />

of symp<strong>to</strong>m development is delayed. Recent<br />

reviews of M. graminicola have been written by<br />

Eyal (1999) and Palmer and Skinner (2002).<br />

17.3.3 Cercospora<br />

This very large form-genus (41000 species)<br />

contains numerous important plant pathogens<br />

associated with a wide range of host plants (Farr<br />

et al., 1989). Examples are C. beticola causing leaf<br />

spot of sugar beet, C. zea-maydis (grey leaf spot of<br />

corn), and C. coffeicola (brown eyespot of coffee).<br />

It is difficult <strong>to</strong> estimate the real number of<br />

species; Johnson and Valleau (1949) isolated<br />

Cercospora from 28 host plants in 16 families,<br />

and all seemed <strong>to</strong> belong <strong>to</strong> the same species.<br />

Further, the dimensions of conidia and conidiophores<br />

can vary in response <strong>to</strong> changes in<br />

humidity. Where known, the teleomorphs of<br />

Cercospora spp. belong <strong>to</strong> Mycosphaerella (Goodwin<br />

et al., 2001). The conidia of Cercospora are


482 LOCULOASCOMYCETES<br />

Fig17.19 Mycosphaerella brassicicola. (a) Section of<br />

a brassica leaf with a subepidermal<br />

pseudothecium containing asci. Note that there is<br />

no hamathecium. (b) Pseudothecium as seen from<br />

above, showing the ostiole. (c) Intact bitunicate<br />

ascus. (d) Stretched ascus resulting from extension<br />

of the endotunica. (e) Ascospores. Scale bar:<br />

40mm (a,b) and10 mm(c e).<br />

hyphomyce<strong>to</strong>us, being produced from pigmented<br />

aerial hyphae. They are long and tapering, as<br />

shown by C. beticola (Fig. 17.21), which is seedborne.<br />

It has long been known that Cercospora<br />

infections are much less severe in shaded<br />

plants as compared <strong>to</strong> those growing in direct<br />

sunlight. The reason for this lies in the production<br />

of a potent pho<strong>to</strong>sensitizing <strong>to</strong>xin, cercosporin<br />

(Fig. 17.22), by many Cercospora spp. A<br />

fascinating and lucid account of cercosporin has<br />

been written by Daub and Ehrenshaft (2000).<br />

Cercosporin, upon absorption of light energy,<br />

is activated <strong>to</strong> an energized state in which it<br />

reacts with molecular oxygen, converting it in<strong>to</strong><br />

various radicals but especially in<strong>to</strong> singlet<br />

oxygen ( 1 O 2 ). This energized form of oxygen<br />

is highly reactive, rapidly destroying organic<br />

molecules and especially membrane lipids.<br />

Cercosporin is a non-selective <strong>to</strong>xin, affecting<br />

bacteria, plants, fungi and animals unless these<br />

produce protective antioxidants such as carotenoids.<br />

It is not yet entirely clear how Cercospora<br />

spp. protect themselves against their own <strong>to</strong>xin,<br />

but part of the answer may lie in keeping it in a<br />

non-reactive (reduced) state. Another mechanism<br />

is the synthesis of vitamin B 6 (pyridoxine) and<br />

its derivatives which can act as antioxidants.<br />

Cercosporin has been found only in Cercospora<br />

species (Goodwin et al., 2001), but biosynthetically<br />

related pho<strong>to</strong>sensitizers are produced by a<br />

range of plant-pathogenic fungi, especially in the<br />

Loculoascomycetes (e.g. Cladosporium, Alternaria,<br />

Stemphylium spp.).<br />

17.3.4 Cladosporium<br />

The anamorph genus Cladosporium is large,<br />

containing about 60 species. A key <strong>to</strong> species in<br />

culture collections has been published by Ho<br />

et al. (1999). Several species have Mycosphaerella<br />

teleomorphs, e.g. C. herbarum is the anamorph<br />

of M. tassiana (von Arx, 1950; Barr, 1958),<br />

C. echinulatum is the anamorph of M. dianthicola<br />

and C. humile the anamorph of M. macrospora. The<br />

association of Cladosporium with Mycosphaerella<br />

has been supported by molecular studies (Crous<br />

et al., 2001). However, most species are without<br />

known teleomorphs. Mycosphaerella tassiana forms


DOTHIDEALES<br />

483<br />

Fig17.20 Wheat leaf blotch symp<strong>to</strong>ms caused by<br />

Mycosphaerella graminicola. (a) Infected leaf showing a necrotic<br />

lesion, in the centre of which pycnidia of Sep<strong>to</strong>ria tritici have<br />

formed (arrows). (b) Septate conidia of S. tritici,morethan<br />

10 times longer than wide.<br />

its pseudothecia on overwintered stalks and<br />

leaves of numerous monocotyledons and dicotyledons<br />

in subarctic and subalpine regions, and a<br />

period of cold is required for ascocarp initiation<br />

in culture (Barr, 1958; Corlett, 1991). In contrast,<br />

C. herbarum is ubiqui<strong>to</strong>us and common in<br />

temperate regions on senescent and dead plant<br />

material, and in soil. Conidia of this and other<br />

Cladosporium spp. are the most abundant component<br />

of the fungal air spora (Gregory, 1973), and<br />

they are probably the most frequent contaminant<br />

of foodstuffs, textiles and paintwork.<br />

They also frequently contaminate cultures of<br />

other fungi in the labora<strong>to</strong>ry. The conidia of<br />

C. herbarum and other common moulds such as<br />

Alternaria alternata and Aspergillus fumigatus are<br />

associated with severe asthma (Zureik et al.,<br />

2002). Over 30 antigens causing mould allergy<br />

have been described from C. herbarum, and most<br />

of them are secre<strong>to</strong>ry or cy<strong>to</strong>plasmic glycoproteins,<br />

often representing common enzymes such<br />

as enolase or aldehyde dehydrogenase<br />

(Breitenbach & Simon-Nobbe, 2002).<br />

Colonies of C. herbarum are dull olive green<br />

<strong>to</strong> black in colour, and appear as a network<br />

of hyphae or a plate-like mass (stroma) of<br />

tightly packed, dark, thick-walled cells (McKemy<br />

& Morgan-Jones, 1991). The conidiophores are<br />

branched or unbranched and conidiogenesis is<br />

holoblastic (Fig. 17.23a). The tip of the conidiophore<br />

bulges out <strong>to</strong> form the first conidium and<br />

it is presumed that all the wall layers of the apex<br />

are involved (Hashmi et al., 1973). The first<br />

conidium buds <strong>to</strong> form a further conidium and<br />

this process continues so that a chain of conidia<br />

develops in acropetal succession, the youngest<br />

conidium at the end of the chain. Most conidia<br />

have a scar (hilum) at each end, but occasionally<br />

a conidium may form two daughter conidia<br />

at its tip so that, as further conidial development<br />

proceeds, a branched chain develops (see<br />

Fig. 17.23a). Such branch-point conidia have been<br />

termed ramoconidia (Lat. ramus ¼ branch) and<br />

are marked by having a single scar at the base<br />

and two scars at the apex. The conidia of<br />

C. herbarum have dark (melanized) walls which<br />

are slightly roughened. They may remain unicellular<br />

or develop 1 3 transverse septa. Another<br />

common species of Cladosporium is C. cladosporioides,<br />

which has smooth conidium walls.<br />

Cladosporium fulvum (also called Fulvia fulva<br />

or Mycovellosiella fulva) is probably not closely<br />

related <strong>to</strong> other Cladosporium spp. No sexual state<br />

has been reported, but other Mycovellosiella<br />

spp., like Cladosporium spp., belong <strong>to</strong><br />

Mycosphaerella (Crous et al., 2001). In contrast <strong>to</strong><br />

most other plant-pathogenic members of the<br />

Loculoascomycetes, C. fulvum is biotrophic. It is<br />

a pathogen of <strong>to</strong>ma<strong>to</strong> plants, especially in<br />

greenhouses. Conidia infect their host through<br />

s<strong>to</strong>mata, and hyphae spread in the leaf apoplast<br />

in close contact with mesophyll cells, but without<br />

producing haus<strong>to</strong>ria. Sucrose, the major<br />

plant transport sugar, is hydrolysed and taken<br />

up, being converted <strong>to</strong> manni<strong>to</strong>l by the fungus<br />

(Joosten et al., 1990). Conspicuous yellow


484 LOCULOASCOMYCETES<br />

incompatible interactions results if the product<br />

of the Avr gene is recognized by the host plant.<br />

Several Avr genes have been characterized, and<br />

their products are usually small proteins<br />

secreted by C. fulvum in<strong>to</strong> the apoplast. Many of<br />

the corresponding Cf genes of <strong>to</strong>ma<strong>to</strong> are also<br />

known; they encode proteins anchored in the<br />

plasma membrane of <strong>to</strong>ma<strong>to</strong> cells, with large<br />

extracellular domains. The examination of the<br />

products of matching avirulence and resistance<br />

genes should provide an opportunity <strong>to</strong> examine<br />

their interactions, and thus the molecular basis<br />

of recognition events involved in specific resistance<br />

(Rivas & Thomas, 2005). This work is still<br />

ongoing.<br />

Fig17.21 Cercospora beticola.Conidiophores and conidia from<br />

sugar beet seed.<br />

(chlorotic) leaf areas are produced as a result of<br />

such systemic infections (Fig. 17.24), and eventually<br />

conidiophores are emitted through<br />

s<strong>to</strong>mata especially on the lower leaf surface,<br />

forming a lawn of spores resembling powdery<br />

mildews but being light brown in colour.<br />

The interaction between C. fulvum and its<br />

<strong>to</strong>ma<strong>to</strong> host is governed by a classical genefor-gene<br />

relationship based on dominant host<br />

resistance genes (Cf genes, for C. fulvum) and<br />

dominant avirulence (Avr) genes in C. fulvum,<br />

i.e. virulence is a recessive trait (Joosten et al.,<br />

1997). The hypersensitive response of<br />

17.3.5 Aureobasidium and black yeasts<br />

Aureobasidium pullulans is a ubiqui<strong>to</strong>us saprotroph<br />

whose main habitats are the phylloplane<br />

and other surfaces of living and senescent plants,<br />

but it can also occur as a symp<strong>to</strong>mless endophyte.<br />

It grows in soil, but is often not recorded<br />

because it is temperature-sensitive and is not<br />

seen on soil plates prepared with warm agar. It<br />

has been isolated from fresh water, estuarine<br />

and marine sediments, sea water, sewage and<br />

other liquid waste (Domsch et al., 1980). Its<br />

teleomorph is Discosphaerina fulvida, a relative of<br />

Mycosphaerella (Yurlova et al., 1999). The fungus<br />

can be readily isolated from leaf washings. It is<br />

pleomorphic, and in culture it forms a rapidly<br />

growing mycelium with wide, septate hyphae<br />

from which intercalary and occasionally terminal<br />

cells give rise <strong>to</strong> single or clustered hyaline<br />

blas<strong>to</strong>conidia by a process of budding. Budding is<br />

associated with local lysis of the wall of the<br />

conidiogenous cell, the inner wall of which then<br />

balloons out and forms the wall of the conidium<br />

(Ramos et al., 1975). Repeated conidium development<br />

from the same or closely adjacent conidiogenous<br />

loci results in the formation of<br />

slimy clusters of hyaline conidia (Fig. 17.25a).<br />

Intercalary cells may enlarge and develop<br />

thicker, dark, melanized walls <strong>to</strong> become<br />

chlamydospores (Fig. 17.25b). Conidia bud in a<br />

yeast-like manner when the fungus is grown<br />

in liquid culture with high inoculum densities.<br />

The yeast cells can continue <strong>to</strong> bud


DOTHIDEALES<br />

485<br />

Fig17.22 Cercosporin.This compound is synthesized as two<br />

polyketide halves which are then fused, accounting for the<br />

symmetric appearance of the molecule.<br />

or may germinate by germ tube. Conidia are<br />

dispersed by rain splash and by air currents. The<br />

fungus may be involved in the biodeterioration<br />

of paint.<br />

Aureobasidium pullulans has a number of<br />

potential applications. It is being investigated<br />

as a possible biocontrol agent against fungi like<br />

Botrytis and Monilia which cause post-harvest<br />

s<strong>to</strong>rage rots of fruit such as grapes, cherries<br />

and strawberries. It is a source of gluconic acid<br />

and of the dextran pullulan, an extracellular<br />

polysaccharide which has uses as an adhesive in<br />

laminates and in fabrics. Pullulan is also used as<br />

Fig17.23 Cladosporium.(a)Cladosporium herbarum; conidiophore with branching chain of blas<strong>to</strong>conidia.The terminal spores<br />

of the chain continue <strong>to</strong> develop blas<strong>to</strong>conidia. (b,c) Cladosporium macrocarpum. (b) Conidiophores and conidia. (c) Conidiophores<br />

developing from a sclerotium.


486 LOCULOASCOMYCETES<br />

Fig17.24 Cladosporium fulvum. Symp<strong>to</strong>ms on<br />

greenhouse-grown <strong>to</strong>ma<strong>to</strong> leaves as seen from above.<br />

The pale leaf areas are chlorotic spots, at the underside of<br />

which a pale greyish-brown felt of conidiophores is produced.<br />

Fig17.25 Aureobasidium pullulans. (a) Blas<strong>to</strong>conidia developing<br />

from an undifferentiated hypha. Several conidia can form<br />

from the same point, i.e. conidiogenesis is polyblastic.<br />

(b) Thick-walled melanized chlamydospores formed from<br />

hyphal segments in an older agar culture.<br />

a low-calorie ingredient of foodstuffs and in<br />

pharmaceutical applications (Deshpande et al.,<br />

1992; Leathers, 2003).<br />

Aureobasidium is a member of a group loosely<br />

called black yeasts. This is not a taxonomic term<br />

but simply describes melanized fungi which<br />

produce yeast states, especially in culture.<br />

Several species of black yeasts, notably species<br />

of Exophiala, Cladophialophora and Ramichloridium,<br />

are known as opportunistic human pathogens<br />

causing infections of the brain and other organs<br />

which can be fatal (Horré & de Hoog, 1999; de<br />

Hoog et al., 2000b). The microscopic features<br />

of these species have been described by de<br />

Hoog (1977) and, <strong>to</strong>gether with supplementary<br />

information, in an excellent compendium by<br />

de Hoog et al. (2000a). Although formerly considered<br />

<strong>to</strong> belong <strong>to</strong> the Loculoascomycetes, they<br />

are now recognized <strong>to</strong> be related <strong>to</strong> Capronia<br />

which is grouped in the Chae<strong>to</strong>thyriales (Haase<br />

et al., 1999; Untereiner, 2000), an order quite<br />

remote from the Loculoascomycetes but with<br />

possible affinities with Plec<strong>to</strong>mycetes or<br />

Lecanorales (Winka et al., 1998). Black yeasts<br />

belonging <strong>to</strong> the Chae<strong>to</strong>thyriales are similar <strong>to</strong><br />

those of the Dothideales not only in microscopic<br />

features, but also in their ecology,<br />

commonly occurring on living and decaying<br />

vegetation.


18<br />

Basidiomycota<br />

18.1 <strong>Introduction</strong><br />

The Basidiomycota (colloquially basidiomycetes)<br />

are a large group of fungi with over 30 000<br />

species. They include many familiar mushrooms<br />

and <strong>to</strong>ads<strong>to</strong>ols, bracket fungi, puffballs, earth<br />

balls, earth stars, stinkhorns, false truffles, jelly<br />

fungi and some less familiar forms. Also classified<br />

here are the rust and smut fungi, which are<br />

pathogens of higher plants and may cause<br />

serious crop diseases. Most basidiomycetes are<br />

terrestrial with wind-dispersed spores, but some<br />

grow in freshwater or marine habitats. Many are<br />

saprotrophic and are involved in litter and wood<br />

decay, but there are also pathogens of trees such<br />

as the honey fungus, Armillaria, which attacks<br />

numerous tree species, and Heterobasidion annosum,<br />

which can seriously damage conifer plantations.<br />

Common woodland mushrooms such as<br />

species of Amanita, Boletus and their allies grow<br />

in a mutually symbiotic relationship with the<br />

roots of trees, forming ec<strong>to</strong>trophic (sheathing)<br />

mycorrhiza. Species of Rhizoc<strong>to</strong>nia, representing<br />

mycelial forms of basidiomycetes, behave as<br />

pathogens <strong>to</strong>wards a wide range of plants but<br />

are mycorrhizal associates of orchids. As saprotrophs,<br />

basidiomycetes play a vital role in<br />

recycling nutrients but they also cause severe<br />

damage as agents of timber decay, e.g. dry rot of<br />

house timbers by Serpula lacrymans. The fruit<br />

bodies (basidiocarps) of many mushrooms are<br />

edible, and some are grown commercially for<br />

food, notably Agaricus bisporus (¼ A. brunnescens,<br />

the white but<strong>to</strong>n mushroom), Pleurotus spp.<br />

(oyster mushrooms) and Lentinula edodes (shiitake).<br />

It is also well known that the basidiocarps<br />

of certain mushrooms are poisonous <strong>to</strong> eat, e.g.<br />

Amanita phalloides (the death cap). Some species<br />

have basidiocarps which are hallucinogenic,<br />

e.g. Amanita muscaria (the fly agaric) and<br />

Psilocybe spp. (‘magic mushrooms’).<br />

The mycelium of basidiomycetes may be very<br />

long-lived. Estimates based on the rate of growth<br />

and the diameter of circles of the fairy ring<br />

fungus Marasmius oreades growing in permanent<br />

pasture show that they may be centuries old. It<br />

has been estimated that the age of an individual<br />

mycelium of Armillaria in a Canadian forest is<br />

at least 1500 years, with an extent of 15 hectares<br />

and a probable biomass in excess of 10 <strong>to</strong>nnes,<br />

making it one of the largest organisms on earth<br />

(Smith et al., 1992).<br />

Not all basidiomycetes grow in the mycelial<br />

form; some are yeast-like and others are dimorphic,<br />

i.e. capable of switching between mycelial<br />

and yeast-like growth. A dimorphic species which<br />

is a dangerous human pathogen <strong>to</strong> immunocompromised<br />

patients is Filobasidiella (¼ Cryp<strong>to</strong>coccus)<br />

neoformans causing cryp<strong>to</strong>coccosis, a fatal disease<br />

of the brain (see pp. 661 664).<br />

18.2 Basidium morphology<br />

The characteristic structure of sexually reproducing<br />

basidiomycetes is the basidium. It is a<br />

spore-bearing cell which produces basidiospores<br />

externally through curved, tapering sterigmata


488 BASIDIOMYCOTA<br />

(Figs. 18.1d,e). Usually there are four spores but in<br />

some cases there are one, two or more than four<br />

basidiospores per basidium. Itersonilia perplexans<br />

has one-spored basidium-like structures (see<br />

Fig. 18.6a), the cultivated mushroom (Agaricus<br />

bisporus) has two-spored basidia, whilst basidia<br />

of the stinkhorn (Phallus impudicus) have as many<br />

as nine spores (see Fig. 20.9b). The form of the<br />

basidium varies, and this has taxonomic significance,<br />

different groups of basidiomycetes having<br />

distinctive types of basidium. In the mushrooms<br />

and their allies, the basidium is a cylindrical cell,<br />

undivided by septa (Fig. 18.1). Such basidia are<br />

termed holobasidia. InDacrymyces and Calocera<br />

(Dacrymycetales) the basidium is undivided by<br />

septa, but the body of the basidium is forked in<strong>to</strong><br />

two, with each arm of the fork developing a single<br />

basidiospore. The basidia of the Jew’s ear fungus,<br />

Auricularia auricula-judae, and related species<br />

are divided by transverse septa, whilst Tremella<br />

and its relatives have basidia with longitudinal<br />

septa. Basidia divided by septa are<br />

termed phragmobasidia or heterobasidia<br />

(Gr. phragmos ¼ a hedge or barricade; heteros ¼<br />

other, different). In rust fungi (Uredinales) and<br />

smut fungi (Ustilaginales) the basidia develop<br />

from thick-walled, originally dikaryotic resting<br />

cells termed teliospores or chlamydospores. A<br />

thin-walled tubular structure, the promycelium,<br />

develops from this resting cell and becomes<br />

divided by transverse septa, each of the resulting<br />

cells producing one basidiospore (in rusts) or<br />

several basidiospores (sporidia) in smuts. Some of<br />

these different kinds of basidia are illustrated in<br />

Fig. 18.2.<br />

18.3 Development of basidia<br />

The development of a holobasidium is readily<br />

observed with the light microscope in the<br />

gill-bearing fungus Oudemansiella radicata which<br />

fruits on dead tree stumps (Figs. 18.1 and 19.18c)<br />

Fig18.1 Oudemansiella radicata. Stages in the development of basidia. (a) Young basidium with numerous vacuoles. Note the clamp<br />

connection at its base and the formation of a further basidial initial. (b) A later stage showing the appearance of a clear apical cap.<br />

(c) Localization of vacuoles <strong>to</strong>wards the base of the basidium. (d) Development of sterigmata and spore initials. A basal vacuole is<br />

enlarging. (e) Fully developed basidium.The spores are full of cy<strong>to</strong>plasm, whilst the body of the basidium contains only a thin lining of<br />

cy<strong>to</strong>plasm surrounding an enlarged vacuole. (f) Discharged basidiospores.


DEVELOPMENT OF BASIDIA<br />

489<br />

Fig18.2 Some different kinds of basidia. (a) Longitudinally divided basidium of Exidiaglandulosa (Tremellales). (b) Tuning-fork type<br />

of basidium of Calocera viscosa (Dacrymycetales). (c) Transversely divided basidium of Auricularia auricula-judae (Auriculariales).<br />

(d) Germinating chlamydospore of Ustilago avenae (Ustilaginales). A transversely septate promycelium has developed and each<br />

segment is forming sporidia. (e) Germinating teliospore of Puccinia graminis (Uredinales). A transversely septate promycelium has<br />

developed and each segment is producing a single basidiospore.<br />

and has particularly large basidia. A detailed<br />

description of the process has been given by<br />

Corner (1948) for O. canarii. Ultrastructural<br />

studies have also been made on a range of<br />

fungi with holobasidia, e.g. the split-gill<br />

(Schizophyllum commune; Wells, 1965), the agarics<br />

Coprinus cinereus (McLaughlin, 1973, 1977, 1982)<br />

and Panellus stypticus (Lingle et al., 1992), the<br />

clavarioid fungus Clavicorona pyxidata (Berbee &<br />

Wells, 1989) and the bolete Boletus rubinellus<br />

(McLaughlin, 1973; Yoon & McLaughlin, 1979,<br />

1984).<br />

18.3.1 Cy<strong>to</strong>logical aspects<br />

In Oudemansiella radicata the basidium arises as<br />

the terminal cell of a hypha making up the gill<br />

tissue on the underside of the cap of the fruit body<br />

(basidiocarp). Basidia are packed tightly <strong>to</strong>gether<br />

in the hymenium at the surface of the gill. A<br />

basidium is at first filled with dense cy<strong>to</strong>plasm,<br />

but soon several small vacuoles appear near its<br />

base. Later these coalesce in<strong>to</strong> a single large<br />

vacuole at the base of the basidium and, by<br />

enlargement of this vacuole, cy<strong>to</strong>plasm is pushed<br />

<strong>to</strong>wards the apex of the basidium. A clear cap is<br />

differentiated at the tip and it is in this region<br />

that the sterigmata develop. Corner (1948) postulated<br />

that there must be four elastic areas in the<br />

upper part of the basidium wall from which<br />

the sterigmata extend. The wall of the upper part<br />

of the basidium consists of two layers, the outer of<br />

which is mucilaginous, the inner firmer. In the<br />

areas where the sterigmata are about <strong>to</strong> bulge<br />

out, a new layer of wall material is deposited<br />

between these two original layers. The outer<br />

mucilaginous layer of the basidial wall bursts<br />

and the apex of the sterigma grows out. It is<br />

surrounded by two wall layers, the inner of which<br />

is continuous with the inner wall layer of the<br />

basidium (Clémençon, 2004). Ultrastructural<br />

studies of Boletus rubinellus show that, in the<br />

region where the sterigmata appear, the basidial


490 BASIDIOMYCOTA<br />

wall is indeed thinner and differs in structure<br />

from the other parts of the wall (Yoon &<br />

McLaughlin, 1984). Beneath these areas there is<br />

some evidence of cy<strong>to</strong>plasmic differentiation,<br />

such as the presence of microtubules and vesicles.<br />

The growth of sterigmata can be compared <strong>to</strong><br />

hyphal tip growth, except that a Spitzenkörper<br />

is absent (McLaughlin, 1973). At the tips of the<br />

sterigmata of Coprinus cinereus, vesicles apparently<br />

fusing with the plasma membrane have been<br />

observed by transmission electron microscopy<br />

studies, and vesicles of similar size were also<br />

found in the basidium (McLaughlin, 1973). Corner<br />

(1948) suggested that the force for the development<br />

of basidia comes from the expansion of the<br />

basal vacuole which acts as a pis<strong>to</strong>n, ramming<br />

the cy<strong>to</strong>plasm in<strong>to</strong> the spores. Ripe basidia thus<br />

contain a large vacuole but very little cy<strong>to</strong>plasm<br />

(Fig. 18.1e). The vacuole, which is filled with<br />

liquid, keeps the basidium turgid until spore<br />

discharge has occurred.<br />

The tip of the sterigma expands <strong>to</strong> form a<br />

small spherical knob, the apophysis (Gr. apo- ¼<br />

away from, separate; physis ¼ growth). Further<br />

development of basidiospores is asymmetric,<br />

expansion being more rapid <strong>to</strong>wards the outside<br />

of the long axis of the basidium. The narrow<br />

point of attachment of the spore at the tip of the<br />

sterigma is the eventual point of spore separation<br />

and is termed the hilum. The non-expanded<br />

part of the apophysis persists as the hilar<br />

appendix (Fig. 18.3a).<br />

18.3.2 Nuclear events<br />

Typically, a basidium is at first binucleate; it is<br />

formed on a dikaryotic mycelium, i.e. a mycelium<br />

with segments containing two haploid nuclei<br />

which are usually genetically different (see<br />

Section 18.9). In this cell nuclear fusion (karyogamy)<br />

occurs (see Fig. 18.4) and is followed immediately<br />

by meiosis, giving rise <strong>to</strong> four haploid<br />

nuclei. As in most fungi, division is intranuclear;<br />

the nuclear membrane remains intact. Meiosis<br />

occurs in the upper part of the basidium. In<br />

narrow basidia the plane of the second meiotic<br />

nuclear division lies parallel <strong>to</strong> the long axis of<br />

the basidium. This type of nuclear division is<br />

termed chiastic (Gr. chias<strong>to</strong>s ¼ crossed, arranged<br />

diagonally) and basidia with nuclear division of<br />

this sort are termed chias<strong>to</strong>basidia. In contrast,<br />

in broader basidia, the plane of both meiotic<br />

nuclear divisions is transverse <strong>to</strong> the long<br />

axis. Nuclear divisions of this type are stichic<br />

(Gr. stichos ¼ a line or row of things) and the<br />

corresponding term for basidia with such division<br />

is stichobasidium. The plane of nuclear<br />

division has relevance <strong>to</strong> taxonomy; chias<strong>to</strong>basidia<br />

are found in mushrooms and <strong>to</strong>ads<strong>to</strong>ols<br />

whilst stichobasidia are more characteristic of<br />

certain genera of bracket fungi in the polyporoid<br />

clade. Terms used <strong>to</strong> define different parts of<br />

a basidium are probasidium, the part within<br />

which karyogamy occurs, and metabasidium,<br />

the part within which meiosis occurs (see Kirk<br />

et al., 2001). The four haploid nuclei formed<br />

during meiosis move in<strong>to</strong> the basidiospores<br />

which are therefore usually four in number. As<br />

they pass through the sterigmata, the nuclei<br />

are often elongated and tapered apically. They<br />

may be led through the sterigmata by microtubules<br />

attached <strong>to</strong> the nuclear spindle pole<br />

bodies (Thielke, 1982; Lingle et al., 1992).<br />

In many basidiomycetes, meiosis is followed<br />

by a post-meiotic mi<strong>to</strong>sis (Duncan & Galbraith,<br />

1972; Clémençon, 2004) which may happen in<br />

different places: (1) In the upper part of the<br />

basidium. Four of the eight nuclei enter the<br />

spores, and those remaining in the basidium<br />

abort. The ripe spores are thus uninucleate, e.g.<br />

in Cantharellus cibarius. (2) At the base of or inside<br />

the sterigmata. One nucleus enters each spore<br />

and the other four remain in the basidium<br />

and degenerate, e.g. in Collybia butyracea. (3) In<br />

the young spore. Four of the daughter nuclei<br />

migrate back in<strong>to</strong> the basidium where they<br />

degenerate, e.g. in Paxillus involutus. (4) In the<br />

young spore, but all eight nuclei remain in the<br />

four spores and none abort; the basidium is left<br />

devoid of nuclei.<br />

18.4 Basidiospore development<br />

On the adaxial side of the apophysis, an electrondense<br />

cy<strong>to</strong>plasmic region appears at the moment


BASIDIOSPORE DEVELOPMENT<br />

491<br />

Fig18.3 Coprinus cinereus.Transmission electron micrographs<br />

of sections illustrating development of basidiospores. (a) Stage<br />

4 basidiospore expanding at the tip of its sterigma.The hilar<br />

appendix body (HAB) is appressed <strong>to</strong> a wall thickening. (b) An<br />

early stage 4 basidiospore with the conical hilar appendix body<br />

in firm contact with the spore plasmalemma.For further<br />

explanation see p. 490. Pho<strong>to</strong>micrographs kindly provided by<br />

D. J. McLaughlin.<br />

of its initiation and, in most cases, persists<br />

throughout the development of the basidiospore,<br />

disappearing only shortly before spore discharge.<br />

It is hemispherical or conical, lying immediately<br />

within the plasma membrane of the apophysis<br />

and closely appressed <strong>to</strong> the wall of the hilar<br />

appendix (see Figs. 18.3a,b). This structure is<br />

the hilar appendix body and has been reported<br />

from several basidiomycetes, including Boletus<br />

rubinellus, Coprinus cinereus (McLaughlin, 1973,<br />

1977), Lactarius lignyotellus (Miller, 1988) and<br />

Panellus stypticus (Lingle et al., 1992). It is probably<br />

present in all ballis<strong>to</strong>sporic species. Its function<br />

is not unders<strong>to</strong>od. Possibly it is involved in the<br />

softening of the closely adjoining wall layers<br />

of the hilar appendix or in the extrusion of<br />

material related <strong>to</strong> the expansion of Buller’s<br />

drop (see Fig. 18.8). Another possibility is that it<br />

blocks the movement of wall vesicles in<strong>to</strong> the<br />

adaxial side of the spore, so possibly contributing<br />

<strong>to</strong> the asymmetric shape of the spore. During<br />

further development, the wall of the hilar<br />

appendix thickens considerably, and uneven<br />

expansion of the wall of the developing spore<br />

occurs. Expansion of the abaxial face is more<br />

rapid, leaving the spore asymmetrically perched<br />

on the sterigma, attached at the hilum and with<br />

the remains of the apophysis forming the hilar<br />

appendix (see Fig. 1.20). In Coprinus cinereus,<br />

McLaughlin (1977) has distinguished four successive<br />

stages of basidiospore expansion (see<br />

Fig. 18.5a).<br />

Stage 1, inception. This is characterized by<br />

the spherical enlargement of the sterigma apex<br />

<strong>to</strong> form a basidiospore primordium 0.6 0.8 mm<br />

in diameter. The hilar appendix body is already<br />

differentiated. The thin basidiospore wall is<br />

three-layered at first. Microtubules are occasionally<br />

present in the sterigma, being orientated<br />

parallel <strong>to</strong> its long axis.<br />

Stage 2, asymmetric growth. The basidiospore<br />

initial grows asymmetrically on its abaxial side,<br />

and the hilar appendix develops. The hilar<br />

appendix body becomes conical and appressed<br />

<strong>to</strong> the plasma membrane of the spore initial. The<br />

hilar appendix is initiated adjacent <strong>to</strong> the hilar<br />

appendix body. The basidiospore wall thickens,<br />

being thickest at the apex of the spore, and is sixlayered.


492 BASIDIOMYCOTA<br />

Fig18.4 Life cycle of the basidiomycete Coprinus (diagrammatic and not <strong>to</strong> scale).The basidiocarp develops from a dikaryotic<br />

secondary mycelium and produces numerous basidia on the surface of its gills beneath the cap. Progressive stages of basidium<br />

maturation involving karyogamy (K) and meiosis (M) are indicated. Eventually, each basidium forms four basidiospores, each<br />

containing a single haploid nucleus. In many basidiomycetes there is a post-meiotic mi<strong>to</strong>sis, giving rise <strong>to</strong> two identical haploid nuclei<br />

in each basidiospore (not shown here). Basidiospores are discharged by the surface-tension catapult mechanism involving Buller’s<br />

drop. Discharged basidiospores germinate <strong>to</strong> form haploid (monokaryotic) mycelia with simple transverse septa. In Coprinus,these<br />

often produce upright conidiophores which form numerous sticky haploid oidia.The apex of a monokaryotic hypha in the vicinity of<br />

an oidium of compatible mating type will respond chemotropically by growing <strong>to</strong>wards the compatible oidium (homing reaction).<br />

Fusion (plasmogamy, P) between the hypha and the oidium initiates the formation of a dikaryotic mycelium bearing clamp<br />

connections. Nuclear fusion does not occur at this stage.The dikaryotic mycelium can develop basidiocarps under appropriate<br />

environmental conditions.Open and closed circles represent haploid nuclei of opposite mating type; the diploid nucleus is drawn<br />

larger and half-filled.<br />

Stage 3, equal enlargement. This is characterized<br />

by spherical enlargement of the basidiospore.<br />

Growth is at an angle of about 45° <strong>to</strong> the<br />

sterigma apex (see Fig. 18.5a). The hilar appendix<br />

body projects further in<strong>to</strong> the spore wall. It is<br />

conical or hemispherical, with the apex of the<br />

cone or base of the hemisphere projecting<br />

<strong>to</strong>wards the hilar appendix. The outermost<br />

layer of the basidiospore wall is sticky.<br />

Stage 4, elongation. Basidiospores grow in<br />

length and a pore cap is formed at the upper end<br />

of the spore. The spore wall becomes darkly<br />

pigmented, starting at the upper end.<br />

Similar asymmetric changes in spore expansion<br />

have been reported by Yoon and McLaughlin<br />

(1984) for basidiospores of Boletus rubinellus<br />

(Fig. 18.5b). Elongation of the spore occurs<br />

mainly at later stages of development.<br />

In certain basidiomycetes, e.g. Amanita<br />

vaginata, the basidiospore is spherical. This<br />

involves even wall expansion during development<br />

and implies that the wall structure is


THE MECHANISM OF BASIDIOSPORE DISCHARGE<br />

493<br />

point of abscission (Yoon & McLaughlin, 1986;<br />

Money, 1998).<br />

18.5 The mechanism of<br />

basidiospore discharge<br />

Fig18.5 Asymmetric expansion of basidiospores during<br />

development. (a) Coprinus cinereus.Changes in shape and the<br />

axis of growth (shown on the right) at successive stages in<br />

basidiospore development.The numbers on the left indicate<br />

spore stages. (b) Boletusrubinellus.Changes in shape and the<br />

major axis of growth (long arrows) at successive<br />

developmental stages.For further explanation see text.<br />

(a)fromMcLaughlin(1977)and(b)fromYoonandMcLaughlin<br />

(1984), by copyright permission of the American Journalof<br />

Botany.<br />

homogeneous. Corner (1948) believed that the<br />

non-spherical shape of basidiospores was the<br />

result of differential setting of the wall material.<br />

This may well be true, but more recent studies<br />

have shown that the basidiospore wall also<br />

varies in thickness. Many basidiospores have<br />

smooth outer walls, but others have characteristic<br />

ornamentations, e.g. spines, folds or ridges.<br />

The ornamentations generally develop by extension<br />

of the outer wall layer of the basidiospore<br />

(Pegler & Young, 1971; Clémençon, 2004). In such<br />

spores, e.g. those of Lactarius and Russula, only<br />

the main body of the spore is ornamented,<br />

whereas the region of the adaxial face immediately<br />

above the hilar appendix is smooth. This<br />

region is termed the suprahilar plage, disc or<br />

depression (Pegler & Young, 1971, 1979). It plays<br />

an important part in the spore discharge<br />

mechanism (see below).<br />

Abscission of the basidiospore from its sterigma<br />

is preceded by the formation of a plug of<br />

material, the hilar plug which blocks the spore<br />

hilum, and a sterigmal plug of wall material<br />

immediately below the hilum. Between the two<br />

plugs a septum appears which represents the<br />

With the exception of the gasteromycetes (see<br />

Chapter 20), in most terrestrial basidiomycetes<br />

basidiospores are ballis<strong>to</strong>spores, i.e. they are<br />

actively projected from basidia. Various suggestions<br />

have been made as <strong>to</strong> the mechanism of<br />

ballis<strong>to</strong>spore discharge (see Webster & Chien,<br />

1990), but the one which we now accept is the<br />

surface tension catapult, originally suggested by<br />

Buller (1922) and Ingold (1939). Discussions of<br />

the various theories of basidiospore discharge<br />

have been written by Webster et al. (1988) and<br />

Money (1998). Shortly before discharge, dissolution<br />

of the abscission layer occurs, indicated by<br />

a slight wobble in the position of the spore. Then<br />

a spherical drop of liquid, Buller’s drop, forms at<br />

the hilar appendix and a shallower liquid<br />

deposit, the adaxial drop (adaxial blob), appears<br />

on the face of the spore above the hilar appendix<br />

(Fig. 18.7). Cinepho<strong>to</strong>graphy has been used <strong>to</strong><br />

illustrate these events (Webster & Hard, 1998b;<br />

Webster, 2006b). Both drops increase in size until<br />

they eventually coalesce, and spore discharge<br />

then immediately occurs (Pringle et al., 2005).<br />

Experimental investigations on the phenomenon<br />

of ballis<strong>to</strong>spore discharge have focused<br />

on Itersonilia perplexans, an unusual heterobasidiomycete<br />

with large ballis<strong>to</strong>spores. This fungus<br />

is a weak plant pathogen and is commonly<br />

associated with lesions caused by other pathogens<br />

such as rust and smut fungi. It also grows in<br />

basidiocarps of certain jelly fungi and can be<br />

readily isolated by allowing it <strong>to</strong> shoot off its<br />

ballis<strong>to</strong>spores from the basidiocarps of Dacrymyces<br />

stillatus or Auricularia auricula-judae (Ingold, 1983a,<br />

1984a). In culture it forms a clamped dikaryotic<br />

mycelium, the tips of whose branches swell <strong>to</strong><br />

form clamped sporogenous cells, each with a<br />

single ballis<strong>to</strong>spore (Fig. 18.6a). The sporogenous<br />

cells do not fully match the definition of basidia<br />

because nuclear fusion and meiosis do not occur<br />

in them; instead the dikaryotic cell forms a


494 BASIDIOMYCOTA<br />

dikaryotic ballis<strong>to</strong>spore directly. A discharged<br />

primary ballis<strong>to</strong>spore may germinate by a germ<br />

tube or by repetition <strong>to</strong> form a secondary<br />

ballis<strong>to</strong>spore (Fig. 18.6b). Yeast-like growth can<br />

also occur, especially on rich media.<br />

Working with Itersonilia, several observations<br />

were made which provided clues <strong>to</strong> the mechanism<br />

of discharge. (1) Ballis<strong>to</strong>spores can be<br />

detached from their sterigmata with a micromanipula<strong>to</strong>r<br />

needle and spores so detached still<br />

develop Buller’s drop. This shows that the liquid<br />

in the drops does not originate from liquid<br />

transported through sterigmata. (2) During<br />

normal discharge, although the volume of<br />

Buller’s drop may attain 60% of that of the<br />

spore, there is no decrease in the dimensions of<br />

the ballis<strong>to</strong>spore, indicating that Buller’s drop<br />

does not come from within the spore. The same<br />

observation has been made on other basidiomycetes<br />

and has led <strong>to</strong> the suggestion that the<br />

liquid in Buller’s drop and also in the adaxial<br />

drop is formed by condensation of water vapour<br />

around a hygroscopic substance extruded from<br />

the hilar appendix and through the spore wall<br />

(Webster et al., 1984a,b, 1989). Washings from the<br />

spores of basidiomycetes belonging <strong>to</strong> several<br />

different taxonomic groups were analysed by<br />

gas liquid chroma<strong>to</strong>graphy, and all gave a positive<br />

result for the presence of manni<strong>to</strong>l. Glucose<br />

was also sometimes detected. The presence of<br />

manni<strong>to</strong>l and hexose in liquid drawn off by<br />

a micropipette from Buller’s drops in Itersonilia<br />

was confirmed by microscope fluorimetry, and<br />

measurements of the solute concentrations in<br />

Buller’s drop corresponded closely <strong>to</strong> the calculated<br />

concentrations which would be necessary<br />

<strong>to</strong> drive the uptake of water from a saturated<br />

atmosphere at the rates observed (Webster et al.,<br />

1995).<br />

The surface tension catapult mechanism<br />

postulates that, as Buller’s drop develops, the<br />

centre of mass of the spore plus drop moves<br />

<strong>to</strong>wards the hilar appendix (Fig. 18.8b). The<br />

coalescence of Buller’s drop with the adaxial<br />

drop causes a rapid redistribution of mass away<br />

from the hilar appendix, resulting in a momentum<br />

which carries the spore plus drop away from<br />

the sterigma (Fig. 18.8c; Webster et al., 1988).<br />

Pringle et al. (2005) have suggested that an even<br />

Fig18.6 Itersonilia perplexans. (a) Basidium (sporogenous cell)<br />

bearing a single ballis<strong>to</strong>spore. Note the clamp connection at<br />

the base of the basidium (arrow). (b) A primary ballis<strong>to</strong>spore<br />

has germinated by repetition <strong>to</strong> form a secondary<br />

ballis<strong>to</strong>spore.<br />

greater momentum may be generated by the<br />

fusion drop moving <strong>to</strong>wards the basidiospore<br />

apex and coming <strong>to</strong> an abrupt halt upon reaching<br />

it. For this mechanism <strong>to</strong> be effective, a rigid<br />

sterigma is required, and it is likely that the<br />

turgor pressure of the vacuolated basidium<br />

contributes <strong>to</strong> the required rigidity (Money, 1998).<br />

The requirement of high humidity for effective<br />

operation of the mechanism is a likely<br />

explanation for observations that basidiospore<br />

concentrations in the air peak at night (Kramer,<br />

1982). High humidity develops in the space<br />

between agaric gills or inside the hymenial<br />

tubes of <strong>to</strong>ads<strong>to</strong>ols and bracket fungi. The<br />

presence of free water would, of course, prevent<br />

operation of the surface tension catapult, and<br />

this may be the reason why agaric basidiocarps<br />

are often umbrella-shaped. The impossibility of<br />

the surface tension catapult mechanism operating<br />

under water also explains why, although<br />

some basidiomycetes grow vegetatively in fresh


NUMBERS OF BASIDIOSPORES<br />

495<br />

Fig18.7 Itersonilia perplexans. Appearance of Buller’s drop and adaxial drop. a 1<br />

a 5<br />

show events immediately preceding and following<br />

the discharge of a secondary ballis<strong>to</strong>spore. An adaxial drop is clearly seen. b 1<br />

b 5<br />

show drops developing on a primary ballis<strong>to</strong>spore.<br />

Note that there is no decrease in the size of the ballis<strong>to</strong>spore as the drops develop. c 1<br />

c 3<br />

illustrate failure of ballis<strong>to</strong>spore discharge.<br />

Here coalescence of Buller’s drop and the adaxial drop is not accompanied by separation of the spore from its sterigma and the<br />

spore soon <strong>to</strong>pples from its perch. Reprinted from Webster et al. (1984a), with permission from Elsevier.<br />

water or the sea, they do not produce ballis<strong>to</strong>spores<br />

there.<br />

The spore of Itersonilia is subjected <strong>to</strong> considerable<br />

acceleration and moves away from the<br />

sterigma at a velocity of over 1 m s 1 . However,<br />

the large surface/mass ratio of a relatively small<br />

object like a ballis<strong>to</strong>spore results in high wind<br />

resistance and therefore rapid deceleration and<br />

loss of momentum, so that the spore soon falls<br />

under the preponderant influence of gravitational<br />

force. The trajec<strong>to</strong>ries of ballis<strong>to</strong>spores<br />

of most fungi follow a short horizontal path<br />

for a distance of about 0.1 0.3 mm and then<br />

turn through a right angle so that spores in still<br />

air drift downwards at a steady sedimentation<br />

velocity. This characteristic trajec<strong>to</strong>ry, termed a<br />

sporabola, ensures that basidiospores projected<br />

in<strong>to</strong> the space between the gills or in<strong>to</strong> the<br />

lumen of a pore turn vertically downwards<br />

before hitting the opposing hymenial surface.<br />

18.6 Numbers of basidiospores<br />

The number of basidiospores produced by a<br />

single basidiocarp can be extremely high. Buller<br />

(1909) estimated that the detached cap of the


496 BASIDIOMYCOTA<br />

mushroom Agaricus campestris produced 1.8 10 9<br />

spores over 2 days at an average rate of<br />

40 million h 1 . Estimates for some other basidiomycetes<br />

are given in Table 18.1. The mycelia of<br />

all these fungi are perennial and an individual<br />

mycelium may produce numerous basidiocarps<br />

over a period of many years. It is therefore clear<br />

that an individual basidiospore has an infinitesimal<br />

chance of successfully establishing a<br />

fruiting mycelium.<br />

Fig18.8 Representation of events associated<br />

with ballis<strong>to</strong>spore discharge. (a) Ballis<strong>to</strong>spore<br />

attached <strong>to</strong> its sterigma before drop<br />

formation.The closed circle within the spore<br />

indicates the centre of mass of the spore.<br />

(b) Buller’s drop appears at the hilar appendix.<br />

The adaxial drop emerges on the spore wall<br />

above it and extends downwards as it<br />

increases in size.The centre of mass of spore<br />

plus drop moves <strong>to</strong> a position nearer the hilar<br />

appendix. (c) Contact between the two drops<br />

is followed by immediate coalescence and the<br />

combined mass of liquid moves rapidly up the<br />

adaxial face of the spore away from the hilar<br />

appendix.The centre of mass moves very<br />

rapidly in the direction of the thin arrow and<br />

the spore drop system gains kinetic energy<br />

and momentum in the same direction,<br />

simultaneously exerting an opposite force F<br />

on the sterigma at the hilum (thick arrow).<br />

Some angular momentum is also exerted on<br />

the spore, related <strong>to</strong> the distance a between<br />

the hilum and the hilar appendix. Reprinted<br />

from Webster et al. (1988), with permission<br />

from Elsevier.<br />

18.7 Basidiospore germination and<br />

hyphal growth<br />

18.7.1 Germination<br />

Basidiospores may remain dormant and retain<br />

viability for several months or even for a few<br />

years if conditions are unsuitable for germination.<br />

Dormancy is frequently exogenous, i.e. the<br />

spores require some external chemical or physical<br />

stimulus before germination can occur.<br />

Germination may be direct by production of a<br />

germ tube, by repetition (i.e. the formation of a<br />

secondary ballis<strong>to</strong>spore), or by the formation of<br />

conidia. Repetitious germination is common in<br />

certain jelly fungi (Tremellales) (see Figs. 18.6b,<br />

21.7b and 21.13d). Germination by the formation<br />

of conidia is illustrated for Dacrymyces stillatus<br />

(Fig. 21.4a), Auricularia auricula-judae (Fig. 21.6c)<br />

and by yeast-like budding in Tremella frondosa (Fig.<br />

21.13c). During direct germination, germ tubes<br />

usually emerge through a special germ pore at<br />

the hilum or, as in Coprinus, through a pore at<br />

the opposite end of the spore.<br />

18.7.2 Monokaryotic and dikaryotic<br />

hyphae<br />

Because the nuclear divisions involved in basidiospore<br />

formation are meiotic, basidiospores<br />

are haploid. Since the post-meiotic nuclear<br />

divisions are mi<strong>to</strong>tic, if there are several nuclei<br />

in a single basidiospore these are usually<br />

genetically identical. They are said <strong>to</strong> be homokaryotic<br />

(Gr. homos ¼ equal, alike; karyon ¼ a nut,<br />

here meaning nucleus). At germination, repeated<br />

mi<strong>to</strong>tic nuclear divisions occur and the early<br />

germ tubes may consequently be multinucleate<br />

and coenocytic. Transverse septa are laid down<br />

behind the growing hyphal tip and eventually<br />

divide the hypha in<strong>to</strong> segments which contain<br />

only a single nucleus. The uninucleate segments<br />

and the hyphae which contain them are said <strong>to</strong><br />

be monokaryotic. The terms homokaryon and<br />

monokaryon or primary mycelium have also<br />

been applied <strong>to</strong> such haploid mycelia. As part of<br />

the sexual cycle, monokaryotic hyphae of genetically<br />

distinct mating type undergo plasmogamy<br />

(soma<strong>to</strong>gamy), i.e. they fuse <strong>to</strong>gether and initiate<br />

the formation of a mycelium made up of


497<br />

BASIDIOSPORE GERMINATION AND HYPHAL GROWTH<br />

Coprinus comatus 5.2 10 9 2days 2.610 9<br />

Table 18.1. Numbers of basidiospores produced from single basidiocarps of selected basidiomycetes. From<br />

Buller (1922).<br />

Species Totalnumber of spores Spore fallperiod Spores discharged perday<br />

Calvatia gigantea 7 10 12<br />

Ganoderma applanatum 5.5 10 12 6months 310 10<br />

Polyporus squamosus 5 10 10 14 days 3.5 10 9<br />

Agaricus campestris 1.6 10 10 6days 2.610 9<br />

segments, each of which contains two genetically<br />

distinct nuclei. Such mycelia are said <strong>to</strong> be<br />

dikaryotic and heterokaryotic. The term secondary<br />

mycelium is also used.<br />

18.7.3 Dolipore septa<br />

The transverse septa which divide both monokaryotic<br />

and dikaryotic hyphae in<strong>to</strong> segments<br />

are incomplete; they contain a central pore<br />

which permits cy<strong>to</strong>plasmic continuity between<br />

adjacent segments. The septal pore is surrounded<br />

by a barrel-shaped flange of thickened wall<br />

material. Such septa, which are characteristic of<br />

basidiomycete mycelia, are known as dolipore<br />

septa (Lat. dolium ¼ large jar, cask). They are<br />

discernible by light microscopy, especially with<br />

ammoniacal Congo red staining, but interpretation<br />

of their structure is only possible in sections<br />

or freeze-fractured cells viewed by electron<br />

microscopy (Figs. 18.9 and 18.10). Septal development<br />

begins by centripetal ingrowth of a<br />

membrane on which wall material (glucan and<br />

chitin) is deposited at both faces from associated<br />

vesicles (for references see Moore, 1985). The<br />

thickening surrounding the pore results from<br />

more rapid deposition of wall material, but there<br />

is evidence that very thick pore rims may be<br />

an artefact associated with chemical fixation.<br />

The pore itself may be blocked by an occlusion<br />

shaped like two champagne corks attached end<br />

<strong>to</strong> end (Fig. 18.9), but blockage of the pore is not<br />

a permanent feature. In some cases there is a<br />

transverse central plate in the pore canal<br />

(Fig. 18.10b).<br />

Overarching the septal pore on each side of<br />

the septum is a specialized portion of endoplasmic<br />

reticulum known as the septal pore cap<br />

or parenthesome (parenthesis ¼ round bracket,<br />

Gr. soma ¼ body). In some cases a second parenthesome<br />

(outer cap) has been reported. In many<br />

basidiomycetes with holobasidia (i.e. Homobasidiomycetes),<br />

the parenthesomes are perforated<br />

(fenestrated), but in many Heterobasidiomycetes,<br />

e.g. Auricularia, there is only a single perforation<br />

or none (Lü & McLaughlin, 1991; Wells, 1994;<br />

Wells & Bandoni, 2001). Other variations in<br />

ultrastructure are known and can be characteristic<br />

of different groups of basidiomycetes, so<br />

that the dolipore/parenthesome complex is<br />

considered <strong>to</strong> be of taxonomic significance<br />

(Khan & Kimbrough, 1982; Moore, 1985, 1996;<br />

McLaughlin et al., 1995; Müller et al., 1998a).<br />

An important role of the dolipore/parenthesome<br />

complex is <strong>to</strong> secure the integrity of hyphal<br />

cells and <strong>to</strong> maintain intercellular communication<br />

and transport of some organelles. A variety<br />

of cy<strong>to</strong>plasmic structures has been reported from<br />

within the pores of Rhizoc<strong>to</strong>nia solani. These<br />

include small tubular and filamen<strong>to</strong>us structures,<br />

small vesicles, tubular endoplasmic reticulum<br />

and other plugging material. The movement<br />

of mi<strong>to</strong>chondria through the septal pore cap<br />

has also been documented (Müller et al., 2000).<br />

Whilst the movement of most organelles<br />

through the septal pore is permitted, the passage<br />

of nuclei is not, and this is possibly a consequence<br />

of their larger size. The migration of<br />

nuclei following plasmogamy between two sexually<br />

compatible monokaryotic mycelia is associated<br />

with enzymatic dissolution of the dolipore<br />

(see below). Another important function of<br />

dolipores is the repair of hyphal damage, the<br />

septal pore being rapidly plugged by electrondense<br />

material in the compartment of a hypha


498 BASIDIOMYCOTA<br />

Fig18.9 Diagrammatic<br />

interpretation of a basidiomycete<br />

dolipore/parenthesome septum.<br />

Reprinted from Moore and<br />

Marchant (1972), by copyright<br />

permission of the National Research<br />

Council of Canada.<br />

adjacent <strong>to</strong> the damaged segment (Aylmore et al.,<br />

1984; Markham, 1994).<br />

18.7.4 Plasmogamy, dikaryotization,<br />

clamp connections<br />

When two compatible monokaryotic mycelia<br />

make contact, the hyphal walls separating<br />

them break down and cy<strong>to</strong>plasmic continuity<br />

(i.e. plasmogamy) of the two monokaryons is<br />

established. In Coprinus cinereus and some other<br />

fungi, plasmogamy may also occur following<br />

fusion of an oidium of one mating type with<br />

a hyphal tip of a compatible monokaryon.<br />

Nuclear migration follows and is associated with<br />

the breakdown of the dolipore/parenthesome<br />

complex <strong>to</strong> permit transfer of a compatible<br />

nucleus from one compartment <strong>to</strong> another<br />

(Giesy & Day, 1965; Marchant & Wessels, 1974).<br />

The mycelium which develops after plasmogamy<br />

is dikaryotic and the process of conversion<br />

of a monokaryon <strong>to</strong> a dikaryon is termed<br />

dikaryotization.<br />

Nuclear fusion (i.e. karyogamy) is delayed<br />

until the basidia have formed, and is thus<br />

preceded by a prolonged dikaryotic state. As a<br />

result of nuclear migration, the tip of a dikaryon<br />

contains two nuclei which are of different<br />

mating types in heterothallic basidiomycetes.<br />

The speed of nuclear migration can be much<br />

higher than the hyphal growth rate. Nuclear<br />

migration rates have been measured in a<br />

number of fungi, e.g. Coprinus cinereus


BASIDIOSPORE GERMINATION AND HYPHAL GROWTH<br />

499<br />

Fig18.10 Longitudinal sections of the dolipore/parenthesome septum in two basidiomycetes as viewed by transmission electron<br />

microscopy. (a) Auriscalpium vulgare, a homobasidiomycete.The parenthesome is perforated. (b) Auricularia auricula-judae,<br />

a heterobasidiomycete.The parenthesome is imperforate. Pho<strong>to</strong>graphs kindly provided by D. J. McLaughlin.<br />

(0.5 1.0 mm h 1 ), Coprinus congregatus (4 cm h 1 )<br />

and Schizophyllum commune (1.5 5.4 mm h 1 )<br />

(Snider, 1965, 1968; Raper, 1966; Ross, 1976).<br />

Microtubules are associated with migrating<br />

nuclei in S. commune (Raudaskoski, 1972) and in<br />

Trametes versicolor (Girbardt, 1968). In the tip cell<br />

and also in the subapical segments (Fig. 18.12) of<br />

a dikaryon, the two nuclei maintain a constant<br />

distance apart from each other, indicating that<br />

they are paired <strong>to</strong>gether by microtubules. They<br />

also move forward <strong>to</strong>gether at a fixed distance<br />

from the apex (Kamada et al., 1993; Torralba et al.,<br />

2004).<br />

The two nuclei in a dikaryotic hyphal tip<br />

divide simultaneously, a process termed<br />

conjugate nuclear division. In most but not all<br />

basidiomycetes, division is accompanied by<br />

nuclear rearrangement involving the formation<br />

of clamp connections, visible as a lateral bulge<br />

in the hyphal wall adjacent <strong>to</strong> a transverse<br />

septum. The events connected with the development<br />

of a clamp connection are set out diagrammatically<br />

in Fig. 18.11. A clamp connection<br />

develops near the position of the pair of nuclei<br />

in the terminal segment of a dikaryotic hypha<br />

(Fig. 18.11b). A backwardly directed hyphal<br />

branch (hook) develops and one daughter<br />

nucleus migrates in<strong>to</strong> it and divides there<br />

mi<strong>to</strong>tically at the same time as mi<strong>to</strong>tic nuclear<br />

division is taking place in the subterminal<br />

nucleus (Fig. 18.11c). A transverse septum develops<br />

in the main hypha between the two daughters<br />

of the subterminal nucleus and an oblique<br />

septum also forms at the base of the hook<br />

(Fig. 18.11d). The hook, containing a single<br />

daughter nucleus, grows round the transverse<br />

septum of the main hypha and its tip fuses with<br />

the wall of the subterminal cell. Plasmogamy<br />

occurs and the nucleus from the hook migrates<br />

in<strong>to</strong> the subterminal cell (Fig. 18.11e). The two<br />

pairs of nuclei then move away from the<br />

transverse septum (Fig. 18.11f). The clamp connection<br />

is a device which ensures that each segment<br />

of a dikaryotic hypha contains two genetically<br />

distinct nuclei. In the absence of clamps or of<br />

some other mechanism for rearrangement of


500 BASIDIOMYCOTA<br />

Fig18.11 Diagrammatic representation of the<br />

sequence of events associated with the formation<br />

of a clamp connection in the dikaryotic hypha of a<br />

basidiomycete. (a) Terminal segment of a hypha<br />

with two genetically dissimilar nuclei. (b) A lateral<br />

bulge in the hypha appears near the paired nuclei.<br />

The leading nucleus moves in<strong>to</strong> the bulge. (c) The<br />

two nuclei undergo conjugate (i.e. simultaneous)<br />

nuclear division. Mi<strong>to</strong>sis of the leading nucleus<br />

occurs within the bulge. (d) The lateral bulge has<br />

developed in<strong>to</strong> a backwardly directed branch or<br />

hook with one daughter of the leading nucleus at<br />

its tip. One of the daughter subterminal nuclei<br />

moves forward.Transverse septa have developed<br />

simultaneously in the main hypha and at the base of<br />

the hook. (e) The tip of the hook has fused with the<br />

wall of the main hypha. Its nucleus has moved in<strong>to</strong><br />

the main hypha and taken up position behind the<br />

daughter of the subterminal nucleus. (f) The pairs<br />

of nuclei move away from the transverse septum in<br />

the main hypha, the terminal pair moving nearer<br />

the hyphal tip and the subterminal pair distally.<br />

Note that both segments contain dissimilar nuclei<br />

but that their arrangement has been reversed.<br />

nuclei, there would be a tendency for dikaryotic<br />

mycelia <strong>to</strong> break down in<strong>to</strong> homokaryotic<br />

segments. Whilst it is reasonable <strong>to</strong> infer that<br />

mycelia with clamps at the septa are dikaryotic,<br />

the converse is not true because there are<br />

numerous basidiomycetes in which the dikaryotic<br />

mycelium does not bear clamps, and this is<br />

especially true of hyphae making up basidiocarps.<br />

For a fuller discussion of dikaryon<br />

formation see Cassel<strong>to</strong>n and Economou (1985).<br />

18.7.5 Aggregates of vegetative hyphae<br />

Basidiomycete hyphae can aggregate <strong>to</strong> form<br />

complex vegetative structures such as hyphal<br />

strands, mycelial cords, mycelial sheets and<br />

rhizomorphs (Butler, 1966; Watkinson, 1979;<br />

Rayner et al., 1985). These often grow more<br />

rapidly than individual hyphae, connect food<br />

bases such as litter fragments in soil, fallen logs<br />

and tree stumps, and are capable of rapid twoway<br />

conduction of water and nutrients.<br />

Mycelial cords<br />

Mycelial cords have been defined by Boddy (1993)<br />

as aggregations of predominantly parallel,<br />

longitudinally aligned hyphae (see Fig. 1.13).<br />

They are especially common in wood-decaying<br />

basidiomycetes but are also involved in the<br />

underground spread of basidiomycetes forming<br />

ec<strong>to</strong>mycorrhiza. Wood-decaying mycelial cord<br />

formers are a very successful ecological group,<br />

using their cords in a variety of strategies <strong>to</strong><br />

secure resources, e.g. by displacing other fungi<br />

from food bases or exploring new resources in<br />

the soil. The cords show relatively little differentiation<br />

in structure and pigmentation.<br />

Examples of fungi with mycelial cords are the<br />

stinkhorn (Phallus impudicus) and the agaric<br />

Megacollybia platyphylla. Their white cords can<br />

extend for many metres in the soil and can be<br />

traced back from the fruit bodies <strong>to</strong> tree stumps<br />

or other food bases if highly motivated students<br />

and digging <strong>to</strong>ols are available (Fig. 18.13a).<br />

Rhizomorphs<br />

The term rhizomorph refers <strong>to</strong> the root-like form<br />

of these mycelial aggregates which are more<br />

highly differentiated than mycelial cords and are<br />

often pigmented brown or black due <strong>to</strong> the<br />

presence of melanin in the walls of small,


ASEXUAL REPRODUCTION<br />

501<br />

Fig18.12 Subapical hypha of a dikaryon of Omphalotus olearius. (a) Interference contrast image showing the segments delimited by<br />

clamp connections (arrowheads). (b) DAPI fluorescence staining of the same hypha showing that each segment contains two paired<br />

nuclei (bright fluorescent objects).<br />

thick-walled cells making up the rind (Townsend,<br />

1954; Cairney, 1991). The best-known example of<br />

a rhizomorph is that of the honey fungus<br />

Armillaria mellea, a serious tree pathogen whose<br />

flat, black, bootlace-like strands often persist<br />

for long periods beneath the bark of trees killed<br />

by the fungus (see pp. 16 18 and Fig. 18.13b).<br />

Many ec<strong>to</strong>mycorrhizal fungi especially in the<br />

bole<strong>to</strong>id clade form hyphal aggregates which<br />

show intermediate features between mycelial<br />

cords and rhizomorphs.<br />

Sclerotia<br />

Sclerotia, an adaptation <strong>to</strong> prolonged survival<br />

and propagation, develop in some basidiomycetes<br />

(see pp. 18 20). They vary in size from<br />

50 mm <strong>to</strong> several centimetres and in weight from<br />

10 mg <strong>to</strong> several kilogrammes. They also vary in<br />

organization from loose aggregations of dark<br />

hyphae <strong>to</strong> highly differentiated structures with a<br />

rind of smaller, dark thick-walled cells and a<br />

medulla of larger, colourless, thin-walled cells<br />

packed with food reserves (Willetts, 1971, 1972;<br />

Clémençon, 2004). Many of the sclerotial types<br />

described on pp. 18 20 are produced by<br />

Basidiomycota, e.g. the loose type (Rhizoc<strong>to</strong>nia<br />

solani) and the strand type (Sclerotium rolfsii). Some<br />

sclerotia are massive, as in Polyporus mylittae<br />

where they may give rise <strong>to</strong> fruit bodies<br />

(carpogenic development; Figs. 18.13c,d).<br />

Basidiocarps of many common agarics such as<br />

Coprinus cinereus, Collybia tuberosa, Hygrophoropsis<br />

aurantiaca and Paxillus involutus may develop from<br />

smaller sclerotia, and these are also characteristic<br />

of the clavarioid fungus Typhula (Corner,<br />

1950). Sclerotial germination, especially of plant<br />

pathogenic fungi such as Rhizoc<strong>to</strong>nia, is more<br />

usually by the outgrowth of mycelium (myceliogenic<br />

germination).<br />

Pseudosclerotia with a similar function <strong>to</strong><br />

sclerotia, but consisting of a compacted mass of<br />

intermixed substratum, soil, s<strong>to</strong>nes, etc., support<br />

the fruiting of certain polypores such as Polyporus<br />

tuberaster (the s<strong>to</strong>ne fungus, tuckahoe), and<br />

Meripilus giganteus.<br />

18.8 Asexual reproduction<br />

Conidium formation is less commonly reported<br />

in the Basidiomycota than in the Ascomycota.<br />

Conidia may develop on monokaryotic or dikaryotic<br />

mycelia, sometimes on both. They may also<br />

form on basidiocarps. Conidia may have an<br />

asexual function in propagation and dispersal<br />

or may also fulfil a sexual role. We can only<br />

consider a few examples. In terms of structure<br />

and on<strong>to</strong>geny, basidiomycete conidia are of three<br />

basic kinds which are summarized below (see<br />

Kendrick & Watling, 1979; Clémençon, 2004).


502 BASIDIOMYCOTA<br />

Fig18.13 Hyphal aggregates of basidiomycetes. (a) White mycelial cords of Megacollybia platyphylla interconnecting a decaying log of<br />

wood serving as food base (far left) with two basidiocarps, exposed by removing the surface leaf litter. (b) Rhizomorphs of Armillaria<br />

mellea in their typical location between the bark (which has been stripped) and the wood of a dead tree. (c,d) The giant<br />

subterranean sclerotium of Polyporus mylittae which has produced an above-ground fruit body (c).When cut open, the sclerotium<br />

consists of a dark rind enclosing a medulla with a texture resembling compacted boiled rice (d). (c,d) reprinted from Fuhrer (2005),<br />

with permission by Bloomings Books Pty Ltd; original images kindly provided by B. Fuhrer.<br />

18.8.1 Oidia (arthroconidia)<br />

The development of arthroconidia in many<br />

ways resembles that found in ascomycetes<br />

(see p. 235). They are often termed oidia. They<br />

may be dry and dispersed by air currents, or wet<br />

and accumulating in slimy masses from which<br />

they are dispersed by insects, in water films or by<br />

rain splash. An example of dry oidia is seen in<br />

the agaric Flammulina velutipes, which fruits<br />

on dead tree stumps and logs in winter (see<br />

Plate 9d). Oidia may develop on monokaryotic<br />

and on clamped, dikaryotic mycelia, but the<br />

oidia formed on dikaryons are monokaryotic.<br />

They are formed by a process of de-dikaryotization<br />

in which each oidium comes <strong>to</strong> contain a<br />

single nucleus (Brodie, 1936; Ingold, 1980).<br />

Vacuoles appear at intervals in aerial hyphae,<br />

separating cylindrical lengths of uninucleate<br />

cy<strong>to</strong>plasm around which cell walls develop. The<br />

wall of the parent hypha then dissolves so that<br />

an irregular fragmented chain of oidia is formed<br />

(Fig. 18.14a). Oidia germinate by germ tubes<br />

formed at either or both ends of the spore. Dry<br />

oidia are not uncommon in members of the<br />

euagarics and polyporoid clades. Wet oidia are<br />

seen in Coprinus cinereus on monokaryons as well<br />

as dikaryons. They are formed on short, erect,<br />

branched or unbranched conidiophores (oidiophores)<br />

(Fig. 18.14b), the tips of which fragment<br />

in<strong>to</strong> uninucleate, smooth-walled, cylindrical


ASEXUAL REPRODUCTION<br />

503<br />

Fig18.14 Arthroconidia in two<br />

Homobasidiomycetes. (a) Arthroconidia of<br />

Flammulina velutipes formed on a monokaryotic<br />

mycelium.When the chains of conidia<br />

disarticulate, the conidia remain dry and are<br />

dispersed by wind currents. (b) Arthroconidia<br />

(oidia) of Coprinus cinereus formed on a<br />

monokaryon.The sticky oidia accumulate in<br />

mucilage in globose heads and are dispersed by<br />

insects visiting the dung on which this fungus<br />

grows. (c) The homing reaction in C. cinereus as<br />

seen in an agar culture.The tips of lateral<br />

branches of monokaryotic hyphae have been<br />

stimulated chemotropically <strong>to</strong> grow or curve<br />

<strong>to</strong>wards oidia of compatible mating type<br />

placed near them a few hours earlier.<br />

Plasmogamy between a hyphal tip and a<br />

compatible oidium is followed by transfer of a<br />

nucleus from the oidium in<strong>to</strong> the monokaryon,<br />

converting it in<strong>to</strong> a dikaryon.<br />

segments (Heinz & Niederpruem, 1970; Polak<br />

et al., 1997). The oidia collect in mucilaginous<br />

globules from which they are dispersed by<br />

insects.<br />

Oidia, whether wet or dry, can function as<br />

spermatizing agents. If an oidium is placed a<br />

little distance ahead of a monokaryotic hypha,<br />

the growing hypha changes direction, being<br />

attracted chemotropically <strong>to</strong>wards the oidium.<br />

This phenomenon is termed the ‘homing reaction’<br />

(Fig. 18.14c). The response has been detected<br />

over distances up <strong>to</strong> 75 mm. This is remarkable in<br />

view of the fact that the width of the approaching<br />

monokaryon hypha is about 2.5 mm and the<br />

growing zone of the hyphal tip is about 0.5 mm<br />

(Kemp, 1975a). The homing reaction is elicited<br />

not only between compatible oidium hypha<br />

combinations but also between incompatible<br />

associations. It may even be triggered by oidia<br />

of different species. Where the oidium and<br />

approaching hypha are compatible, plasmogamy,<br />

i.e. fusion of the hyphal tip and the<br />

oidium, takes place, followed by nuclear<br />

migration and the eventual establishment of a<br />

dikaryon. Plasmogamy may also occur between<br />

an oidium and an unrelated approaching hypha,<br />

i.e. one belonging <strong>to</strong> a different species. In this<br />

case the introduction of a nucleus from the<br />

oidium in<strong>to</strong> a cell of the unrelated hypha results<br />

in a lethal response, involving the death of the<br />

recep<strong>to</strong>r cell and possibly some adjacent cells.<br />

Kemp (1975a,b) has argued that the lethal<br />

response is important in maintaining interspecific<br />

barriers.<br />

18.8.2 Blastic conidia<br />

Blastic development involves the marked<br />

enlargement of a recognizable conidium initial<br />

before the conidium is delimited by a septum.<br />

This is usually achieved by a blowing-out of part<br />

of the conidiophore wall. There are many<br />

different ways in which blastic conidia can<br />

develop (Kendrick & Watling, 1979). For example,<br />

they may develop singly, synchronously in<br />

clusters, or in succession. The growth form


504 BASIDIOMYCOTA<br />

Auricularia auricula-judae, a parasite of elder<br />

(Sambucus nigra), develops successive blastic<br />

conidia from conidiophores on germinating<br />

basidiospores (Fig. 21.6c) or from monokaryotic<br />

hyphae. At discharge, each basidiospore is<br />

unicellular, but three transverse septa divide<br />

the basidiospore before germination and each of<br />

the resulting cells may form a short conidiophore<br />

which swells and curves at the tip <strong>to</strong> form<br />

a horseshoe-shaped (lunate) conidium, followed<br />

by further similar conidia.<br />

Sis<strong>to</strong>trema hamatum (anamorph Ingoldiella<br />

hamata) is a subtropical aquatic basidiomycete<br />

with large, septate, branched conidia (Fig. 25.17;<br />

Nawawi & Webster, 1982). They consist of a main<br />

axis over 400 mm long with two <strong>to</strong> three tapering<br />

laterals. The tips of the branches are recurved.<br />

Conidia are formed on dikaryotic and on<br />

monokaryotic mycelia. Dikaryotic conidia are<br />

distinguished by the presence of clamp<br />

connections at the septa (Figs. 25.17a,b) which<br />

are absent from the otherwise similar monokaryotic<br />

conidia (Fig. 25.17c). Other species<br />

with aquatic branched conidia are recognizable<br />

as basidiomycetes either by the presence of<br />

clamp connections or dolipore septa within<br />

their conidia and mycelia (Nawawi, 1985;<br />

Webster, 1992).<br />

Fig18.15 Spiniger state of Heterobasidion annosum.<br />

Conidiophores and conidia.Conidial development is blastic.<br />

adopted by many basidiomyce<strong>to</strong>us yeasts is an<br />

example of blastic development (see p. 659).<br />

Heterobasidion annosum, a tree-pathogenic<br />

polypore, forms clusters of dry blas<strong>to</strong>conidia<br />

synchronously on the swollen tips of upright,<br />

club-shaped conidiophores (Fig. 18.15). Following<br />

detachment of the conidia, the surface of the<br />

swollen tip of the conidiophore bears spiny<br />

denticles. These conidiophores resemble the<br />

Oedocephalum type of conidia found in certain<br />

ascomycetes (see Fig. 14.4), but the anamorph<br />

name Spiniger has been given <strong>to</strong> them (Stalpers,<br />

1974).<br />

18.8.3 Chlamydospores<br />

The term chlamydospore is used here in a wide<br />

sense following the definition by Kirk et al. (2001)<br />

as ‘an asexual one-celled spore (primarily for<br />

perennation, not dissemination) originating<br />

endogenously and singly within part of a preexisting<br />

cell, by the contraction of the pro<strong>to</strong>plast<br />

and possessing an inner secondary and often<br />

thickened hyaline or brown wall, usually impregnated<br />

with hydrophobic material.’ A common<br />

example of a basidiomycete forming chlamydospores<br />

is Laetiporus sulphureus, a yellowish-orange<br />

bracket fungus parasitic on a range of tree hosts<br />

such as oak, willow and yew. In culture they are<br />

mostly formed terminally on aerial branched<br />

conidiophores which develop from a mycelium<br />

lacking clamp connections (Fig. 18.16a). In the<br />

mycelium within the substrate, intercalary chlamydospores<br />

are also formed. Adjacent <strong>to</strong> the


ASEXUAL REPRODUCTION<br />

505<br />

Fig18.16 Sporotrichum state of Laetiporus sulphureus. (a) Branched aerial conidiophores with terminal chlamydospores.<br />

(b) Intercalary chlamydospores. (c) Detached chlamydospores. (a) and (b) <strong>to</strong> same scale.<br />

Fig18.17 SEM image of a bulbil of Minimedusa polyspora<br />

produced in agar culture.<br />

swollen segment packed with cy<strong>to</strong>plasm which<br />

becomes the chlamydospore, there are empty<br />

mycelial segments (Fig. 18.16b). The collapse of<br />

these empty cells brings about release of the<br />

chlamydospore (Fig. 18.16c) so that release is<br />

rhexolytic, as already seen in certain ascomycete<br />

conidia (see p. 235).<br />

Fig18.18 Bulbillomyces farinosus.Fragment of a large<br />

multicellular, branched propagule made up of inflated cells<br />

with clamp connections at their base. Air is trapped between<br />

the cells. From Abdullah (1980), with permission.<br />

18.8.4 Bulbils<br />

Certain basidiomycetes develop multicellular,<br />

pseudoparenchyma<strong>to</strong>us propagules composed<br />

of thin-walled, undifferentiated, homogeneous<br />

cells in the form of raspberry-like bulbils. A<br />

classical example is the terrestrial basidiomycete<br />

Minimedusa polyspora (Fig. 18.17; Weresub &<br />

LeClair, 1971). Another good example of a multicellar<br />

propagule is seen in the semi-aquatic


506 BASIDIOMYCOTA<br />

fungus Bulbillomyces farinosus (anamorph Aegerita<br />

candida). The anamorphic state (Fig. 18.18)<br />

appears on the surface of wet wood from freshwater<br />

streams as white clusters of clamped cells<br />

between which air is trapped, giving the propagule<br />

buoyancy. This is a typical feature of aeroaquatic<br />

fungi which form asexual propagules<br />

in air on the surface of leaves and branches of<br />

trees previously submerged in fresh water (see<br />

Section 25.3).<br />

18.9 Mating systems in<br />

basidiomycetes<br />

18.9.1 Homothallic systems<br />

About 10% of the Basidiomycota which have been<br />

tested are homothallic (Raper, 1966). Three types<br />

of homothallic behaviour may be distinguished,<br />

namely primary, secondary and unclassified<br />

homothallism.<br />

Primary homothallism<br />

In Coprinus sterquilinus a single basidiospore<br />

germinates <strong>to</strong> form a mycelium, which soon<br />

becomes organized in<strong>to</strong> binucleate segments<br />

bearing clamp connections at the septa. There<br />

is no genetic distinction between the two nuclei<br />

in each cell, and this mycelium is capable of<br />

forming fruit bodies.<br />

Secondary homothallism<br />

(pseudohomothallism)<br />

In Coprinus ephemerus f. bisporus the basidia<br />

bear only two spores, but the spores are heterokaryotic.<br />

After meiosis two nuclei enter each<br />

spore and a mi<strong>to</strong>tic division may follow. On<br />

germination, a single spore forms a dikaryotic<br />

mycelium with clamp connections, capable of<br />

fruiting. Occasional spores, on germination, give<br />

rise <strong>to</strong> non-clamped mycelia, and fruiting occurs<br />

only if these are paired in certain combinations,<br />

showing that the fungus is basically heterothallic.<br />

The cultivated mushroom, Agaricus bisporus,<br />

also has a mating system of this type. Although<br />

most basidia bear two spores, four-spored basidia<br />

do occur, and when monosporous cultures<br />

derived from basidiospores from four-spored<br />

basidia are crossed, they produce fruiting mycelia<br />

in certain combinations. It has been suggested<br />

that a simple bipolar system (see below) is<br />

operating (Miller, 1971; Raper et al., 1972).<br />

This situation occurs in a number of other twospored<br />

basidiomycetes and is closely paralleled<br />

by that found in certain four-spored ascomycetes<br />

such as Neurospora tetrasperma (Raper, 1966).<br />

<strong>Fungi</strong> showing both secondary homothallism<br />

and heterothallic behaviour are said <strong>to</strong> be<br />

amphithallic.<br />

Unclassified homothallism<br />

The four-spored wild mushroom, Agaricus campestris,<br />

is homothallic in the sense that a mycelium<br />

derived from a single spore is capable of fruiting.<br />

There is nuclear fusion in the basidium, followed<br />

by two nuclear divisions, presumably meiotic.<br />

However, paired nuclei, conjugate nuclear divisions<br />

and clamp connections have not been<br />

observed.<br />

18.9.2 Heterothallic systems<br />

Amongst the remaining 90% of the<br />

Basidiomycota reported <strong>to</strong> be heterothallic, we<br />

can distinguish bipolar and tetrapolar<br />

conditions.<br />

Bipolar<br />

In species such as Coprinus comatus (the shaggy<br />

ink-cap) and Pip<strong>to</strong>porus betulinus (the birch polypore),<br />

when mycelia obtained from single spores<br />

from any one fruit body are mated <strong>to</strong>gether,<br />

dikaryons are formed in half the crosses. This can<br />

be explained on the basis of a single gene<br />

(or fac<strong>to</strong>r) with two alleles. Because only a<br />

single fac<strong>to</strong>r is involved, the genetic basis for<br />

the bipolar condition is described as unifac<strong>to</strong>rial.<br />

Segregation of the two alleles at meiosis ensures<br />

that a single spore carries only one allele.<br />

Dikaryons are only formed between monokaryons<br />

carrying different alleles at the mating<br />

type locus. In fact, it is known that there may<br />

be numerous mating type alleles in a population<br />

of fruit bodies collected over a wide area (see<br />

below). About 25% of Basidiomycota examined<br />

have been shown <strong>to</strong> be bipolar. Most members of<br />

the Uredinales and Ustilaginales have mating


MATING SYSTEMS IN BASIDIOMYCETES<br />

507<br />

systems of this type, although a few have more<br />

complex systems.<br />

Tetrapolar<br />

In the coprophilous ink-cap Coprinus cinereus or<br />

the wood-rotting Schizophyllum commune, fertile<br />

dikaryons result in one-quarter of the matings<br />

when primary mycelia derived from basidiospores<br />

from a single fruit body are intercrossed.<br />

The explanation originally proposed for this<br />

situation was that incompatibility is controlled<br />

by two genes (fac<strong>to</strong>rs), with two alleles at each<br />

locus. Because two separate fac<strong>to</strong>rs are involved,<br />

the genetic basis is termed bifac<strong>to</strong>rial. Thus we<br />

can denote the two genes as A and B and their<br />

two alleles as A 1 , A 2 and B 1 , B 2 , respectively.<br />

Consider the cross of a monokaryon bearing A 1 B 1<br />

with another bearing A 2 B 2 . This would result in<br />

a fertile dikaryon (A 1 B 1 þ A 2 B 2 ). Such a dikaryon<br />

would form spores following meiosis and the<br />

spores would be of four kinds: A 1 B 1 , A 2 B 2<br />

(parentals), A 2 B 1 and A 1 B 2 (recombinants). In<br />

most cases studied, the proportions of the four<br />

kinds of spore are equal, showing that the A<br />

and B loci are unlinked, i.e. borne on different<br />

chromosomes.<br />

Fertile dikaryons are only formed when the<br />

alleles present at each locus in the opposing<br />

monokaryons differ e.g. in crosses of the type<br />

A 1 B 1 A 2 B 2 or A 2 B 1 A 1 B 2 . Where there is an<br />

identical allele at either or both loci the cross<br />

is unsuccessful. Thus the success of inbreeding<br />

within the spores of any one fruit body is only<br />

25% in tetrapolar species as compared with 50%<br />

in bipolar forms.<br />

A species with tetrapolar heterothallism<br />

whose life cycle is unusual and difficult <strong>to</strong><br />

interpret is Armillaria mellea. Most of the cells of<br />

the mycelium are monokaryotic, and there is no<br />

evidence of clamp connections in the mycelium<br />

or the rhizomorphs. Fruit body primordia arise<br />

from the monokaryotic rhizomorphs, and the<br />

cells of the young primordia are also monokaryotic.<br />

However, cells making up the gill tissue<br />

are dikaryotic, and these dikaryotic hyphae are<br />

associated with clamp connections, whilst the<br />

monokaryotic cells formed in the remaining<br />

tissue of the stem and cap have no clamps.<br />

Estimations of nuclear volume in monokaryotic<br />

and dikaryotic cells suggest that the nuclei of<br />

monokaryotic cells are diploid, whilst those of<br />

dikaryotic cells are haploid. It is presumed that<br />

the diploid nuclei undergo haploidization by an<br />

unknown mechanism during the formation of<br />

gill initials. Within the basidia, nuclear fusion<br />

and meiosis occur, and a single meiotic product<br />

enters each basidiospore. In the spore, the<br />

nucleus divides mi<strong>to</strong>tically, and one daughter<br />

nucleus from each spore migrates back in<strong>to</strong> the<br />

body of the basidium and degenerates (Korhonen<br />

& Hintikka, 1974; Tommerup & Broadbent, 1974;<br />

Ullrich & Anderson, 1978; Anderson & Ullrich,<br />

1982).<br />

Variations in the life cycle of A. mellea have<br />

been reported. In a form designated as ‘Japanese<br />

A. mellea’, Ota et al. (1998) presented evidence that<br />

four haploid nuclei appear after meiotic division<br />

of the diploid nucleus in the young basidium.<br />

These haploid nuclei fuse in pairs, resulting in<br />

two diploid nuclei which migrate in<strong>to</strong> two of<br />

the developing basidiospores where they divide<br />

mi<strong>to</strong>tically. One nucleus from each basidiospore<br />

returns <strong>to</strong> the basidium, leaving the spore<br />

containing one diploid nucleus. Occasionally<br />

nuclear migration fails <strong>to</strong> occur and the<br />

spore remains binucleate. Spores with diploid<br />

nuclei can complete the life-cycle by forming<br />

a mycelium competent <strong>to</strong> fruit. Ota et al. (1998)<br />

therefore concluded that the ‘Japanese<br />

A. mellea’ illustrates a kind of secondary<br />

homothallism.<br />

18.9.3 Multiple alleles, complex loci<br />

Although a single spore from one fruit body of<br />

S. commune or C. cinereus is compatible with only<br />

one-quarter of its fellow spores, crosses between<br />

spores from fruit bodies of different origin often<br />

result in 100% mating success, i.e. a spore from<br />

one fruit body can mate successfully with 100%<br />

of the spores from a different fruit body. The<br />

explanation of this phenomenon is that a large<br />

number of alleles is present in a population<br />

representing the species as a whole, instead of<br />

the single pair of alleles at each locus present<br />

in any one dikaryotic mycelium. Suppose that a<br />

second fruit body had the composition (A 3 B 3 þ<br />

A 4 B 4 ), then all the four kinds of spore it produced,


508 BASIDIOMYCOTA<br />

A 3 B 3 , A 3 B 4 , A 4 B 3 and A 4 B 4 would be compatible<br />

with all the spores of the original fruit body, on<br />

the assumption that the essential requirement<br />

for fertility is that in any cross both alleles<br />

should differ at both loci. This high value for<br />

outbreeding success implies the existence of a<br />

large number of different mating type fac<strong>to</strong>rs,<br />

and estimates based on isolates from worldwide<br />

collections of fruit bodies indicate that the<br />

number of mating type fac<strong>to</strong>rs of certain species<br />

may be many thousands (Raper, 1966).<br />

Our understanding of the structure and<br />

function of the mating type fac<strong>to</strong>rs is derived<br />

from studies of three species, namely the two<br />

Homobasidiomycetes S. commune and C. cinereus,<br />

and the maize smut fungus Ustilago maydis<br />

(Ustilaginomycetes). Extensive literature is available<br />

on general aspects of this <strong>to</strong>pic (see Kües &<br />

Cassel<strong>to</strong>n, 1992; Kämper et al., 1994; Kothe, 1996;<br />

Kronstad & Staben, 1997; Brown & Cassel<strong>to</strong>n,<br />

2001; Cassel<strong>to</strong>n, 2002) and on the individual<br />

species C. cinereus (Kües, 2000), S. commune<br />

(Stankis et al., 1990; Ullrich et al., 1991; Kothe,<br />

1999) and U. maydis (Banuett, 1995).<br />

18.9.4 Functions of the A and B loci<br />

There is a close similarity between the functions<br />

and structure of the A and B mating type loci<br />

in S. commune and C. cinereus as described in<br />

Table 18.2. The fact that they influence<br />

many different functions indicates that their<br />

gene products are active as regula<strong>to</strong>ry proteins.<br />

The A locus encodes two peptides which <strong>to</strong>gether<br />

make up a heterodimer transcription fac<strong>to</strong>r,<br />

i.e. the molecule is active only if its two halves<br />

are different from each other (see below and<br />

Fig. 23.8). The B locus of S. commune and C. cinereus<br />

directly encodes the peptide pheromone and a<br />

transmembrane recep<strong>to</strong>r for pheromones of<br />

compatible strains. In this way it differs from<br />

the mating system of the ascomycete yeast<br />

S. cerevisiae in which the mating type loci<br />

encode regula<strong>to</strong>ry genes whose products stimulate<br />

the transcription of pheromone and pheromone<br />

recep<strong>to</strong>r genes located elsewhere in the<br />

genome (see Fig. 10.5).<br />

18.9.5 Structure of themating type fac<strong>to</strong>rs<br />

The results of crossing experiments between<br />

compatible monokaryons indicate that recombination<br />

may occur within the A and B loci<br />

<strong>to</strong> give novel mating types. This implies that<br />

both loci are complex. For S. commune it has<br />

been proposed that the A locus contains two subloci,<br />

Aa and Ab. Similarly the B locus contains two<br />

sub-loci, Ba and Bb. For each sub-locus, pairing<br />

tests revealed a number of ‘alleles’: 9 Aa, 32Ab,<br />

9 Ba and 9Bb (Ullrich et al., 1991). Recombination<br />

between the Aa and Ab ‘alleles’ gives rise <strong>to</strong> 288<br />

(i.e. 9 32) different A specificities. When these<br />

are multiplied by the 81 B specificities, over<br />

20 000 possible mating types are generated. In<br />

C. cinereus there are an estimated 160 A specificities<br />

and 79 B specificities which <strong>to</strong>gether<br />

generate over 12 000 mating types (Cassel<strong>to</strong>n,<br />

2002). The actual number of Aa and Ab alleles is<br />

not known (Cassel<strong>to</strong>n & Olesnicky, 1998).<br />

Table18.2. Functions of the A and B loci in Schizophyllum commune and Coprinus cinereus. The functions operate<br />

only if there are different specificities at the A and B loci.<br />

Locus<br />

A-regulated<br />

B-regulated<br />

Function<br />

Pairing of nucleiin dikaryon<br />

Initiation of clamp cell formation<br />

Synchronized nuclear division<br />

Septation<br />

Nuclear exchange between monokaryons<br />

Septal dissolution and nuclear migration<br />

Peg formation and clamp cell fusion<br />

Pheromone production


MATING SYSTEMS IN BASIDIOMYCETES<br />

509<br />

In both S. commune and C. cinereus it has been<br />

possible <strong>to</strong> obtain details of the structure of the<br />

sub-loci at a finer level of resolution. For<br />

example, the Aa locus of C. cinereus contains one<br />

gene pair (designated a1 and a2) and the Ab locus<br />

two gene pairs (b1 and b2, d1 and d2), each with a<br />

number of alleles. In the field, many strains have<br />

lost one or more of their maximum complement<br />

of six A genes (Fig. 18.19). The gene products<br />

are the subunits of the heterodimer transcription<br />

fac<strong>to</strong>r. The two different subunits (1 and 2)<br />

are encoded by compatible alleles. For example,<br />

a functional heterodimer can be formed from<br />

the product of an a allele at the a1 position<br />

(e.g. a1 1) and a different one at the a2 position<br />

(e.g. a2 2, a2 3, etc., but not a2 1). Likewise,<br />

heterodimers can be formed between the<br />

products of two compatible b or two d alleles,<br />

but there is a great deal of functional redundancy<br />

in the system in the sense that it is<br />

sufficient if only one of six possible heterodimers<br />

is formed. In other words, compatibility between<br />

two strains at the A locus is ensured if compatible<br />

alleles are present at the a or b or d genes.<br />

Cassel<strong>to</strong>n and Olesnicky (1998) have calculated<br />

that only 5 6 alleles would be required at each<br />

gene pair (e.g. 5 6 6) <strong>to</strong> account for the<br />

estimated 160 unique A gene combinations in<br />

C. cinereus.<br />

The genetics of mating type behaviour in the<br />

maize smut fungus Ustilago maydis is, in many<br />

respects, similar <strong>to</strong> that described for C. cinereus<br />

and S. commune. This fungus is dimorphic, with<br />

a monokaryotic, saprotrophic yeast-like state<br />

which can be readily cultured, and a dikaryotic<br />

mycelial state which is parasitic on maize<br />

and requires living host cells for growth (see<br />

Fig. 23.1). Ustilago maydis is tetrapolar, unlike<br />

most smut fungi which are bipolar. Incompatibility<br />

is governed by two loci, designated<br />

a and b. Unfortunately the functions of the<br />

b mating type fac<strong>to</strong>r in U. maydis correspond <strong>to</strong><br />

those of the A fac<strong>to</strong>r in C. cinereus and S. commune,<br />

and vice versa. There are two alleles at the a locus<br />

(a 1 and a 2 ) and about 25, possibly more, at the b<br />

locus. The a locus encodes a pheromone and a<br />

pheromone recep<strong>to</strong>r; it controls the switch <strong>to</strong><br />

filamen<strong>to</strong>us growth but not pathogenicity. The b<br />

locus encodes the production of a heterodimeric<br />

DNA-binding protein involved in self/non-self<br />

recognition and also in pathogenicity; two<br />

different b alleles are required for pathogenic<br />

growth (see p. 643).<br />

Much lower numbers of alleles have been<br />

estimated for the bird’s nest fungi (Gasteromycetes;<br />

see p. 581). For instance, Cyathus striatus<br />

has 4 A and 5 B alleles, and Crucibulum vulgare 3 A<br />

and about 16 B alleles. It has been suggested that<br />

Fig18.19 The A loci of two different field strains of Coprinus cinereus. Neither strain has the complete set of six genes, but none the<br />

less four different functional heterodimers canbe formed from gene products of different (compatible) alleles at the a, b and d genes.<br />

Only one would be sufficient for overall compatibility at the A locus. Note that the position of the d genes is reversed relative <strong>to</strong><br />

those of the a and b genes.Genetic recombination events are possible in homologous DNA regions (thick lines) but not within the<br />

regions covered by the gene pairs a, b or d. Redrawn and modified from Cassel<strong>to</strong>n and Olesnicky (1998).


510 BASIDIOMYCOTA<br />

the small number of incompatibility fac<strong>to</strong>rs may<br />

be related <strong>to</strong> the specialized method of dispersal<br />

in which numerous basidiospores are packaged<br />

in<strong>to</strong> a peridiole.<br />

The molecular basis of bipolar (unifac<strong>to</strong>rial)<br />

mating systems may be quite similar if one<br />

considers that here the A and B loci are simply<br />

linked <strong>to</strong> each other in close proximity on the<br />

same chromosome (see p. 637).<br />

18.9.6 The Buller phenomenon<br />

Buller (1931) discovered that a monokaryon of<br />

C. cinereus paired with a dikaryon of the same<br />

fungus could be converted <strong>to</strong> the dikaryotic<br />

state. The same phenomenon has been reported<br />

in some other bipolar and tetrapolar fungi.<br />

Conversion, i.e. dikaryotization, is brought<br />

about by nuclear migration from the dikaryon<br />

in<strong>to</strong> the monokaryon. Different kinds of combination<br />

(di mon matings) are possible (Raper,<br />

1966).<br />

There are two kinds of legitimate combinations.<br />

(1) In fully compatible combinations, a<br />

monokaryon is compatible with both nuclear<br />

components of the dikaryon, e.g. bipolar (A 1 þ<br />

A 2 ) A 3 or tetrapolar (A 1 B 1 þ A 2 B 2 ) A 3 B 3 . (2) In<br />

hemicompatible combinations, a monokaryon is<br />

compatible with only one of the nuclear components<br />

of the dikaryon, e.g. bipolar (A 1 þ A 2 ) A 2<br />

or tetrapolar (A 1 B 1 þ A 2 B 2 ) A 1 B 1 .Inillegitimate<br />

(incompatible) combinations, a monokaryon is<br />

compatible with neither nuclear component of<br />

the dikaryon, e.g. tetrapolar (A 1 B 1 þ A 2 B 2 ) A 1 B 2<br />

or A 2 B 1 .<br />

Surprising features were discovered in some<br />

such pairings. In compatible pairings using<br />

Schizophyllum it was found that the selection of<br />

a compatible nucleus from the dikaryon <strong>to</strong><br />

dikaryotize the monokaryon was not a matter<br />

of chance. Consider the fully compatible di mon<br />

mating (A 1 B 1 þ A 2 B 2 ) A 3 B 3 . If conversion of<br />

the monokaryon by one of the nuclei from<br />

the dikaryon were entirely random, dikaryons<br />

(A 1 B 1 þ A 3 B 3 ) and (A 2 B 2 þ A 3 B 3 ) would be equally<br />

frequent. However, this is not the case; there is<br />

evidence of preferential selection of one mating<br />

type over the other, but the reasons for this<br />

selection are obscure. A second unexpected<br />

feature is the discovery that dikaryotization can<br />

occur in incompatible pairings. A possible reason<br />

for this phenomenon is that somatic recombination<br />

between the nuclei in the original dikaryon<br />

can occur <strong>to</strong> give rise <strong>to</strong> a nucleus compatible<br />

with that of the monokaryon (Raper, 1966).<br />

An unusal mating phenomenon which is the<br />

equivalent of the Buller phenomenon has been<br />

discovered in Armillaria mellea. The vegetative<br />

phase of this fungus is mainly diploid, not<br />

dikaryotic. Its mating system is bifac<strong>to</strong>rial, i.e.<br />

tetrapolar, controlled by A and B loci. When<br />

diploid and haploid mycelia are paired in certain<br />

combinations, mating occurs, i.e. the diploid<br />

mycelium is capable of dikaryotizing the haploid<br />

monokaryon (Anderson & Ullrich, 1982).<br />

18.10 Fungal individualism:<br />

vegetative incompatibility<br />

between dikaryons<br />

When genetically distinct dikaryons belonging <strong>to</strong><br />

the same species are paired <strong>to</strong>gether, they do not<br />

coalesce. Although their hyphae may fuse<br />

<strong>to</strong>gether, the cells of the resulting heteroplasmon<br />

die and often become darkly pigmented.<br />

This is the result of vegetative incompatibility. In<br />

contrast, when genetically identical dikaryons<br />

are paired, their mycelia intermingle. Vegetative<br />

incompatibility is readily demonstrated in<br />

culture (Fig. 18.20c) and is recognizable in the<br />

field as black bands of fungal cells at the interface<br />

between adjacent dikaryotic colonies in<br />

decaying tree stumps (Fig. 18.20a). The phenomenon<br />

was first discovered in the bracket fungus<br />

Trametes versicolor, the cause of white-rot in<br />

deciduous trees, but has since been found <strong>to</strong> be<br />

widespread amongst different ecological groups<br />

of basidiomycetes, including wood rotting, coprophilous,<br />

ec<strong>to</strong>mycorrhizal and plant pathogenic<br />

species, and has led <strong>to</strong> the concept of fungal<br />

individualism (see Todd & Rayner, 1980; Rayner,<br />

1991a,b). Pairing tests between dikaryotic<br />

isolates facilitate the determination of the<br />

limits and extent of an individual mycelium<br />

of a basidiomycete. In T. versicolor, tree trunks


RELATIONSHIPS<br />

511<br />

colonized by this fungus show a number of decay<br />

columns which can be traced through a series of<br />

transverse slices of wood extending the length<br />

of the trunk. In stumps of birch (Betula) these<br />

columns can be matched genetically with different<br />

clusters of fruit bodies emerging at the<br />

surface of the stump.<br />

By making pairings of isolates from basidiocarps,<br />

decaying wood, and mycelial cords or<br />

rhizomorphs of Megacollybia platyphylla or Armillaria<br />

bulbosa, it has been shown that a single<br />

individual can extend over many hectares<br />

(Anderson et al., 1979; Thompson & Rayner,<br />

1982). Isolations from basidiocarps of the fairy<br />

ring fungus Marasmius oreades show that the<br />

same individual mycelium has spread in an<br />

annular fashion and estimates of the incremental<br />

growth rate indicate that the same individual<br />

may be several centuries old (Mallett & Harrison,<br />

1988; Dix & Webster, 1995).<br />

Vegetative incompatibility enables a fungus <strong>to</strong><br />

distinguish between ‘self’ and ‘non-self’ and<br />

prevents the spread of genetic information in the<br />

form of nuclei and mi<strong>to</strong>chondria and possibly also<br />

of fungal viruses from one mycelium <strong>to</strong> another<br />

of the same species. It thus helps <strong>to</strong> preserve the<br />

genetic integrity of an individual mycelium.<br />

18.11 Relationships<br />

Basidiomycota are related <strong>to</strong> Ascomycota.<br />

Evidence for this view is based on similarities<br />

in composition and construction of the hyphal<br />

wall (e.g. presence of chitin; see Section 1.2.2),<br />

the molecular basis of mating type control<br />

(see p. 266), the production of similar conidial<br />

states, and molecular sequence data (see Bruns<br />

et al., 1992; Tehler et al., 2003). It has been<br />

Fig18.20 Vegetative incompatibility inTrametes versicolor. (a) Section of a tree stump containing several different dikaryotic<br />

colonies. At the interfaces between adjacent colonies double black lines indicate the incompatibility reaction. (b,c). Interactions<br />

between dikaryotic colonies in culture. (b) Reaction when two genetically identical colonies are inoculated near each other.<br />

The two mycelia intermingle. (c) Incompatible reaction when two genetically distinct dikaryons are paired.


512 BASIDIOMYCOTA<br />

Fig18.21 Phylogenetic tree of the<br />

Basidiomycota based on small<br />

nuclear (18S) rDNA gene analyses.<br />

The various taxonomic groups and<br />

the chapters covering them are<br />

indicated. Redrawn and modified<br />

from Nishida and Sugiyama (1994),<br />

with permission from Mycoscience.<br />

postulated that, in evolutionary terms, basidiospores<br />

are the equivalent of ascospores whose<br />

development has become external instead of<br />

taking place endogenously. Clamp connections<br />

(Fig. 18.11), characteristic of dikaryotic hyphae<br />

of basidiomycetes, are seen as homologous <strong>to</strong> the<br />

croziers in ascogenous hyphae (see Fig. 8.10);<br />

both have the same function of re-distributing<br />

nuclei. Tehler et al. (2003) have indicated the<br />

strength of their belief that the two groups are<br />

closely related by classifying them <strong>to</strong>gether in<br />

the Dikaryomycota. They have suggested that<br />

this group is a sister group <strong>to</strong> the Glomeromycota<br />

(treated in this book within the Zygomycota<br />

as the order Glomales, see Section 7.6). Ascomycetes<br />

and basidiomycetes probably diverged from<br />

zygomycetes some 400 600 million years ago<br />

(Berbee & Taylor, 2001).<br />

Basidiomycetes have a long his<strong>to</strong>ry. Fossil<br />

records show that the characteristic feature of<br />

basidiomycete hyphae, the clamp connection,<br />

already existed some 300 million years ago in<br />

the Carboniferous period (Dennis, 1970). Clearly<br />

recognizable mushroom basidiocarps have been<br />

preserved in amber about 90 94 million years<br />

old (Hibbett et al., 1997a). This also contained<br />

well-preserved basidiospores with prominent<br />

hilar appendices.<br />

18.12 Classification<br />

The classification which we have chosen follows<br />

that proposed by McLaughlin et al. (2001). They<br />

have divided the phylum Basidiomycota in<strong>to</strong><br />

four classes:<br />

1. Homobasidiomycetes. <strong>Fungi</strong> with holobasidia,<br />

e.g. agarics and polypores.<br />

2. Heterobasidiomycetes. <strong>Fungi</strong> with heterobasidia,<br />

i.e. jelly fungi and their allies.<br />

3. Urediniomycetes. Rust fungi.<br />

4. Ustilaginomycetes. Smut fungi.


CLASSIFICATION<br />

513<br />

These taxonomic groups are broadly<br />

supported by phylogenetic analyses (Swann &<br />

Taylor, 1993). An example of the kind of arrangement<br />

currently proposed for the basidiomycetes<br />

is shown in Fig. 18.21. Because their basidia are<br />

arranged on a freely exposed hymenium, the<br />

Heterobasidiomycetes and Homobasidiomycetes<br />

are sometimes collectively termed Hymenomycetes.<br />

Various basidiomycetes in which the<br />

hymenium is not freely exposed and which do<br />

not discharge their basidiospores violently have<br />

adopted alternative methods of spore dispersal.<br />

They are related <strong>to</strong> a number of Homobasidiomycete<br />

groups, from which they have arisen in<br />

the course of evolution. Whilst recognizing that<br />

they do not form a natural group, they are here<br />

considered <strong>to</strong>gether as gasteromycetes because<br />

they share important biological features (see<br />

Chapter 20). Basidiomyce<strong>to</strong>us yeasts are another<br />

artificial assemblage of fungi which have relationships<br />

with various taxonomic groups; these<br />

are also considered in a separate chapter<br />

(Chapter 24).


19<br />

Homobasidiomycetes<br />

19.1 <strong>Introduction</strong><br />

<strong>Fungi</strong> included in the Homobasidiomycetes<br />

possess holobasidia, in contrast <strong>to</strong> the heterobasidia<br />

(phragmobasidia) of the<br />

Heterobasidiomycetes, rusts and smuts<br />

(see Chapters 21 23). The traditional classification<br />

of the Homobasidiomycetes, founded by the<br />

nineteenth century Swedish mycologist Elias<br />

Fries, was based on a number of different<br />

arrangements of the hymenium on the hymenophore.<br />

The most common types, shown in<br />

Fig. 19.1, are (A) agaricoid, i.e. gill-bearing<br />

(lamellate); (B) poroid, i.e. bearing pores instead<br />

of gills; (C) hydnoid, i.e. with a <strong>to</strong>othed or spiny<br />

hymenium; (D) clavate, with a club-shaped<br />

or coralloid fruit body, the outside of which<br />

is covered by the hymenium; (E) resupinate,<br />

i.e. with a flattened (corticioid) hymenium<br />

appressed <strong>to</strong> the underside of solid surfaces;<br />

and (F) epigeous or (G) hypogeous gasteroid or<br />

secotioid (non-ballis<strong>to</strong>sporic) hymenophores<br />

(see Chapter 20). It has long been suspected<br />

that the different hymenial arrangements have<br />

evolved separately in unrelated fungal groups,<br />

i.e. that they represent examples of convergent<br />

evolution. They can be interpreted as different<br />

ways of maximizing the hymenial area for a<br />

given amount of fungal tissue (Pöder, 1983;<br />

Pöder & Kirchmair, 1995). Examples of similar<br />

hymenophore arrangements in unrelated fungi<br />

are seen in the tubular hymenia characteristic<br />

of Boletus (Fig. 19.21) and Trametes (Fig. 19.26)<br />

or the <strong>to</strong>othed hymenia found in the<br />

homobasidiomycete Hydnum (Fig. 19.1c) and the<br />

heterobasidiomycete Pseudohydnum (Fig. 21.9b).<br />

Quite possibly, only one or a few genes are<br />

involved in the morphogenetic events resulting<br />

in these various hymenial structures, as has been<br />

shown for the transition of gill-bearing fungi <strong>to</strong><br />

gasteromycetes (see pp. 578 580). There is much<br />

other evidence, e.g. in fruit body construction,<br />

ultrastructure, chemical reactions of basidiocarps,<br />

basidium cy<strong>to</strong>logy, spore colour and<br />

morphology, and molecular sequence data <strong>to</strong><br />

support the view that gross morphological<br />

characters of hymenophores are unsatisfac<strong>to</strong>ry<br />

criteria on which <strong>to</strong> base a natural classification<br />

system.<br />

Hibbett et al. (1997b) and Hibbett and Thorn<br />

(2001), in their preliminary attempts at a<br />

more natural classification, found the<br />

Homobasidiomycetes <strong>to</strong> be distributed amongst<br />

eight phylogenetic clades (Fig. 19.2), and this<br />

scheme has been confirmed and extended in<br />

subsequent work (e.g. Binder & Hibbett, 2002;<br />

Moncalvo et al., 2002). The eight clades may<br />

ultimately be given the status of orders, and<br />

although it is tempting <strong>to</strong> use some of the<br />

existing order names in synonymy, e.g.<br />

Agaricales for the euagarics clade, the necessary<br />

emendations of orders have not yet been<br />

carried out. Therefore, we shall use the clade<br />

system for the time being. One serious limitation<br />

of it is that all large-scale attempts at<br />

Homobasidiomycete phylogeny undertaken <strong>to</strong><br />

date have been based on nuclear and mi<strong>to</strong>chondrial<br />

ribosomal DNA, and confirmation by<br />

comparing other genes will be required before


INTRODUCTION<br />

515<br />

Fig19.1 Examples of hymenial surfaces in the Homobasidiomycetes. (a) Gill-bearing (lamelloid) surface of Agaricus silvaticus.<br />

(b) Tubular (poroid) hymenium of Boletus badius.(c)Spiny(hydnoid)surfaceofHydnumrepandum. (d) Club-shaped (clavate) fruit body<br />

of Clavariadelphus pistillaris.The hymenium lines the surface of the fruit body. (e) Flattened (corticioid) hymenium of Peniophora<br />

quercina forming a crust on the underside of an oak twig. (f) Enclosed (gasteroid) fruit body of Scleroderma citrinum.<br />

a more definitive taxonomic system can be put in<br />

place.<br />

The eight clades, shown in Fig. 19.2, are as<br />

follows: (1) the polyporoid clade, including most<br />

members of the former order Polyporales; (2) the<br />

euagarics clade containing many members of<br />

the old order Agaricales, <strong>to</strong>gether with fungi<br />

from diverse other groups; (3) the bole<strong>to</strong>id clade;<br />

(4) a thelephoroid clade; (5) the russuloid clade<br />

including a particularly wide range of fruit body<br />

types; (6) the hymenochae<strong>to</strong>id clade; (7) the<br />

cantharelloid clade; and (8) a gomphoid phalloid<br />

clade. Table 19.1 indicates that each of these<br />

clades contains several of the hymenophore


Fig19.2 The eight-clade phylogenetic system of the Homobasidiomycetes based on both nuclear and mi<strong>to</strong>chondrial small-subunit<br />

rDNA sequences.Gasteromycete genera are printed in bold. Redrawn from Hibbett et al.(1997b),withpermission.<br />

ß 1997 National Academy of Sciences, U.S.A.


STRUCTURE AND MORPHOGENESIS OF BASIDIOCARPS<br />

517<br />

Table 19.1. Distribution of fruit body morphotypes among the major clades of Homobasidiomycetes shown<br />

in Fig. 19.2. From Hibbett and Thorn (2001).<br />

Major clades<br />

Hymenophore type<br />

(A)<br />

Agaricoid<br />

(B)<br />

Poroid<br />

(C)<br />

Hydnoid<br />

(D)<br />

Clavate<br />

(E)<br />

Resupinate<br />

(F)<br />

Epigeous<br />

gasteroid<br />

secotioid<br />

(G)<br />

Hypogeous<br />

gasteroid<br />

(1) Polyporoid þ þ þ þ þ þ<br />

(2) Euagarics þ þ þ þ þ þ<br />

(3) Bole<strong>to</strong>id þ þ þ þ þ þ<br />

(4) Thelephoroid þ þ þ þ þ<br />

(5) Russuloid þ þ þ þ þ þ þ<br />

(6) Hymenochae<strong>to</strong>id þ þ þ þ þ<br />

(7) Cantharelloid þ þ þ þ<br />

(8) Gomphoid phalloid þ þ þ þ þ þ<br />

morphotypes shown in Fig. 19.1. This finding<br />

is perhaps best explained by the assumption that<br />

the first Homobasidiomycetes had a morphologically<br />

simple fruit body form, probably the<br />

flattened (resupinate) type, from which more<br />

complex fruit bodies arose as numerous independent<br />

evolutionary events (Hibbett & Binder,<br />

2002). The class Gasteromycetes is an artificial<br />

taxon, comprising gasteroid and secotioid fruit<br />

body forms which have lost the capacity for<br />

violent basidiospore discharge and have adopted<br />

alternative mechanisms for spore dispersal<br />

(see Fig. 19.2). Because of unifying biological<br />

and morphological features, the gasteromycetes<br />

are described separately in Chapter 20, where<br />

we will refer <strong>to</strong> their affinity <strong>to</strong> the<br />

Homobasidiomycete clades.<br />

19.2 Structure and morphogenesis<br />

of basidiocarps<br />

The fleshy basidiocarps of many fungi belonging<br />

<strong>to</strong> groups 2 5, 7 and 8 in Fig. 19.2 (‘agarics’)<br />

differ from the <strong>to</strong>ugher fruit bodies found in the<br />

polyporoid and hymenochae<strong>to</strong>id clades 1 and 6<br />

(‘polypores’) in texture and construction. Agaric<br />

basidiocarps are often umbrella-shaped, with a<br />

centrally attached stalk (stipe) supporting the<br />

cap (pileus) beneath which are the gills or tubes<br />

lined by basidia. The polypores include bracket<br />

fungi with basidiocarps which often bear pores<br />

and are laterally attached <strong>to</strong> their substratum.<br />

These fruit bodies are usually firmer in texture,<br />

i.e. leathery, corky or woody. The differences in<br />

texture reflect fundamental principles of<br />

construction as shown by analysis of the<br />

hyphae composing the basidiocarp. There are<br />

also differences in development between agaricand<br />

polypore-type fruit bodies.<br />

19.2.1 Hyphal analysis<br />

Corner (1932a,b, 1953) dissected the basidiocarp<br />

tissues of various polypores and showed that<br />

they were constructed of three distinctive types<br />

of hyphae generative, binding and skeletal<br />

(see below). Later work by Corner and others has<br />

identified a further range of hyphal types<br />

making up basidiocarps (see Pegler, 1996; Kirk<br />

et al., 2001; Clémençon, 2004). This method of<br />

hyphal analysis provides valuable criteria which<br />

have greatly improved the taxonomy of the<br />

genera of agaric- and polypore-type fungi.<br />

However, it does not of itself provide the basis<br />

for a higher-level ‘natural classification’ of<br />

homobasidiomycetes.


518 HOMOBASIDIOMYCETES<br />

There are three main hyphal types. (1)<br />

Generative hyphae are thin-walled near the<br />

margin of a basidiocarp but often thickerwalled<br />

behind, with or without clamp connections,<br />

usually with cy<strong>to</strong>plasmic contents. This<br />

kind of hypha is universally present in all<br />

basidiocarps at some stage of development.<br />

The generative hyphae produce basidia and<br />

other types of cell making up the hymenium,<br />

and they also give rise <strong>to</strong> the other kinds<br />

of hyphae from which the basidiocarp is<br />

constructed (Fig. 19.3a). (2) Skeletal hyphae are<br />

unbranched or sparsely branched, thick-walled<br />

hyphae with a narrow lumen. They arise as<br />

lateral branches of generative hyphae and form<br />

a rigid framework (Fig. 19.3c). (3) Binding<br />

hyphae (sometimes termed ligative hyphae) are<br />

much-branched, narrow, thick-walled hyphae of<br />

limited growth. These hyphae weave themselves<br />

between the other hyphae and bind them<br />

<strong>to</strong>gether (Fig. 19.3b).<br />

Several other kinds of hypha have been<br />

described, some as intermediates between the<br />

above three principal systems. Sarco-hyphae<br />

are composed of long, greatly inflated, mostly<br />

unbranched cells 500 3000 10 30 mm<br />

with relatively narrow septa. They can be interpreted<br />

as skeletal, inflated generative hyphae.<br />

In Amauroderma rugosum (Ganodermataceae)<br />

skele<strong>to</strong>-ligative hyphae resembling skeletal<br />

hyphae have thick-walled con<strong>to</strong>rted branches<br />

and function in the same way as binding<br />

hyphae. Arboriform skeletal hyphae with terminal<br />

thick-walled branches are present in the<br />

basidiocarps of Ganoderma (see Figs. 19.23b,c).<br />

Gloeoplerous hyphae have dense oily contents<br />

(Fig. 19.27). The diverse hyphal types may be<br />

present in basidiocarps in different<br />

Fig19.3 Hyphal analysis of material dissected<br />

from the fruit body of a trimitic polypore,<br />

Trametes versicolor. (a) Generative hyphae<br />

characterized by thin walls, dense cy<strong>to</strong>plasmic<br />

contents and clamp connections. (b) Binding<br />

hyphae, branched, con<strong>to</strong>rted and thick-walled.<br />

The arrow shows the origin from a generative<br />

hypha. (c) A skeletal hypha, unbranched and<br />

thick-walled, originating from a generative<br />

hypha (arrow).


STRUCTURE AND MORPHOGENESIS OF BASIDIOCARPS<br />

519<br />

combinations according <strong>to</strong> the mitic system of<br />

description (Gr. mi<strong>to</strong>s ¼ a thread of the warp).<br />

Monomitic basidiocarps are made up of<br />

generative hyphae only. Most agaric fruit bodies<br />

are of this type, containing inflated generative<br />

hyphae with or without clamps. However,<br />

various modified cell types may also be present,<br />

e.g. the lactifers (laticifers) containing latex in<br />

Lactarius (Fig. 19.4). In Lactarius and Russula<br />

(russuloid clade), the flesh contains rosettes<br />

of globose, thin-walled sphaerocysts or sphaerocytes<br />

(see Figs. 19.4 and 19.9c) which give it a<br />

brittle texture. Basidiocarps of the polypore<br />

Bjerkandera adusta are also monomitic, but<br />

here the walls of the generative hyphae thicken<br />

with age. Sarcomitic construction is seen in<br />

the polypore-type basidiocarps of Meripilus, made<br />

up of inflated sarco-hyphae which function as<br />

skeletal elements.<br />

Dimitic fruit bodies, produced by various<br />

members of the polyporoid and russuloid clades,<br />

may show several kinds of construction. Dimitic<br />

fruit bodies with binding hyphae (i.e. generative<br />

and binding hyphae) are found in<br />

the basidiocarp of Laetiporus sulphureus (see<br />

Plate 10b). The dissepiments (i.e. the blocks of<br />

flesh separating the tubes) are, however, monomitic.<br />

Dimitic fruit bodies with skeletal<br />

hyphae are found in Heterobasidion annosum<br />

(see Fig. 19.26b), whereas dimitic basidiocarps<br />

with skele<strong>to</strong>-ligative hyphae are found in Lentinus<br />

and Ganoderma.<br />

Trimitic fruit bodies contain all three kinds<br />

of hypha shown in Fig. 19.3. A good example<br />

is Trametes versicolor, a common bracket fungus<br />

on hardwood stumps, trunks and branches<br />

(see Plate 10a). At the growing margin of<br />

the fruit body and in the dissepiments, construction<br />

is dimitic, composed of generative and<br />

skeletal hyphae. Binding hyphae develop in the<br />

adult flesh behind the growing margin. The term<br />

sarcotrimitic has been used <strong>to</strong> describe fruit<br />

body construction in Trogia where there are<br />

generative, binding and sarco-hyphae.<br />

19.2.2 Development of basidiocarps<br />

Basidiocarps begin their development from a<br />

hyphal knot, an aggregation of hyphae formed<br />

usually on the secondary (i.e. dikaryotic) mycelium.<br />

The surface of the hyphae forming the<br />

fruit body primordium is often non-wettable<br />

due <strong>to</strong> hydrophobin rodlets (Wessels, 1994, 2000).<br />

This property ensures that air-filled channels<br />

are present within basidiocarps, allowing efficient<br />

gas exchange <strong>to</strong> occur even under wet<br />

conditions (Lugones et al., 1999). The first-formed<br />

hyphae making up the young fruit body are<br />

little differentiated from normal vegetative<br />

hyphae and are termed protenchyma but, as<br />

differentiation proceeds, the hyphae of the<br />

Fig19.4 Lactariusrufus. Cells from the pileus,<br />

including thin-walled septate generative hyphae,<br />

a wider, thicker-walled laticiferous hypha which<br />

contains a milky latex, and clusters of globose<br />

sphaerocysts.


520 HOMOBASIDIOMYCETES<br />

basidiocarp become interwoven and often tightly<br />

packed, making up the plectenchyma. If the<br />

hyphae are less tightly packed, the resulting<br />

tissue is known as pseudoparenchyma. Although<br />

a range of cell types can be distinguished in the<br />

mature basidiocarp, they all develop from<br />

generative hyphae. Enlargement and differentiation<br />

of basidiocarps is associated with inflation<br />

of segments of these hyphae. Details of the<br />

morphogenesis of basidiocarps will be described<br />

later for selected genera.<br />

Several kinds of development have been<br />

described, relating <strong>to</strong> whether or not the<br />

hymenophore is at all times exposed or is at<br />

first surrounded by other tissues. For example,<br />

young fruit bodies may be enveloped by a<br />

universal veil which is broken as the pileus<br />

expands, leaving a cup-like volva at the base of<br />

the stipe and broken scales on the cap as in<br />

Amanita. In some agarics the hymenophore is<br />

protected during development by a partial veil<br />

stretching from the edge of the cap <strong>to</strong> the stem.<br />

Where the partial veil is thin and cobweb-like as<br />

in Cortinarius it is termed the cortina, but where<br />

it is composed of firmer tissues it persists as a<br />

ring (annulus) on the stem. Agaric fruit bodies<br />

may thus have both universal and partial veils<br />

(Fig. 19.5), either, or neither. These different<br />

kinds of development have been distinguished by<br />

technical terms as follows (Reijnders, 1963, 1986;<br />

Moore, 1998).<br />

Gymnocarpic hymenophores are naked from<br />

the time of their first appearance and are never<br />

enclosed by tissue. The pileus develops at the<br />

tip of the stipe and the hymenophore differentiates<br />

on the lower side. Gymnocarpic development<br />

is found in several unrelated genera, e.g.<br />

Cantharellus, Boletus, Russula, Lactarius and Cli<strong>to</strong>cybe<br />

(Fig. 19.6a).<br />

The angiocarpic hymenophore is enclosed by<br />

tissue during part of its development. There are<br />

two kinds of angiocarpic development. In primary<br />

angiocarpy, the pileus margin, hymenophore<br />

and sometimes also pileus and stipe<br />

differentiate beneath the surface of the primary<br />

tissue of the primordium (protenchyma).<br />

Stropharia semiglobata and Amanita rubescens are<br />

primarily angiocarpic (Figs. 19.6c,d). In a fruit<br />

body showing secondary angiocarpy, hyphae<br />

from an already differentiated surface grow out<br />

<strong>to</strong>wards the exterior <strong>to</strong> enclose the primordium<br />

or part of it. The hyphae may extend from the<br />

margin of the pileus <strong>to</strong>wards the stipe, or from<br />

the stipe <strong>to</strong> the pileus, or both. In Lentinus tigrinus<br />

hyphae from both the pileus margin and the<br />

stipe extend <strong>to</strong> enclose the developing gills<br />

(Fig. 19.6b).<br />

The tissue of the monomitic agaric-type<br />

basidiocarp is plectenchyma<strong>to</strong>us or pseudoparenchyma<strong>to</strong>us.<br />

The fruit body expands due <strong>to</strong><br />

inflation of the cells. Although differentiation<br />

in<strong>to</strong> skeletal or binding hyphae is limited<br />

Fig19.5 Schematic drawing of an<br />

agaric-type fruit body showing both<br />

the universal and partial veils. (a) The<br />

but<strong>to</strong>n stage. (b) Fully expanded fruit<br />

body. Remnants of the universal veil<br />

are seen as cap scales and the volva,<br />

whereasthepartialveilhasformeda<br />

ring (annulus) around the stipe.


STRUCTURE AND MORPHOGENESIS OF BASIDIOCARPS<br />

521<br />

Fig19.6 Basidiocarp developmentin some Agaricales illustratedby longitudinal sections (after Reijnders,1963). (a) Cli<strong>to</strong>cybeclavipes.<br />

Gymnocarpic development. (b) Lentinus tigrinus. Secondary angiocarpy resulting from extension of hyphae from pileus margin and<br />

stipe <strong>to</strong> enclose the previously differentiated hymenophore. (c) Stropharia semiglobata. Primary angiocarpy. Note the universal veil<br />

enclosing the upper part of the primordium. In mature fruit bodies this becomes gelatinous.The hymenophore is also enclosed by a<br />

partial veil. (d) Amanitarubescens.Tangential section. Note the break up of the universal veil <strong>to</strong> form scales on the surface of the<br />

pileus.The gill chamber is enclosed by a partial veil.<br />

or absent, specialized tissues or cells may arise<br />

(see Fig. 19.4). In a study of the fine structure of<br />

the sporophore of Agaricus campestris, Manocha<br />

(1965) has shown that the stipe contains two<br />

kinds of cells wide inflated cells and narrower<br />

thread-like cells. A similar differentiation is<br />

found in Coprinus cinereus (Moore, 1998). When<br />

portions of stipe tissue are placed on suitable


522 HOMOBASIDIOMYCETES<br />

agar media, only the thinner hyphae seem <strong>to</strong><br />

give rise <strong>to</strong> vegetative growth (Borriss, 1934).<br />

19.2.3 Types of lamellae<br />

The gills of most lamellate fruit bodies are<br />

wedge-shaped in longitudinal section and are of<br />

the aequi-hymenial (aequi-hymeniiferous) type.<br />

This term refers <strong>to</strong> the fact that the hymenium<br />

develops in an equal manner all over the surface<br />

of the gill, i.e. basidial development is not<br />

localized at any one point on the gill. The<br />

wedge-shaped section may be an adaptation <strong>to</strong><br />

minimize wastage of spores should the fruit body<br />

be tilted from the vertical. Buller (1909) calculated<br />

for the field mushroom Agaricus campestris<br />

that a displacement of 2°30’ from the vertical<br />

would still allow all the spores <strong>to</strong> escape.<br />

Adjustments in the orientation of the stipe and,<br />

sometimes, of the gills themselves may further<br />

help <strong>to</strong> minimize wastage. The <strong>to</strong>pic of gravitropism<br />

(gravimorphogenesis), i.e. the re-orientation<br />

by stipe bending or gill curvature in<br />

displaced basidiocarps <strong>to</strong> res<strong>to</strong>re hymenia <strong>to</strong> a<br />

vertical position, is discussed on p. 546.<br />

Gills of the inaequi-hymenial (inaequihymeniiferous)<br />

type are characteristic of the<br />

ink-caps (Coprinus sensu la<strong>to</strong>) where the gills are<br />

not wedge-shaped in section, but parallel-sided,<br />

and often held apart by cystidia (see Fig. 19.7).<br />

The term inaequi-hymenial refers <strong>to</strong> the fact that<br />

the hymenium develops in an unequal manner,<br />

with basidia ripening in zones. In Coprinus a wave<br />

Fig19.7 Coprinus atramentarius.Vertical section of the parallel-sided gills showing basidia, interspersed by globose paraphyses<br />

and a cystidium extending across the space between adjacent gills <strong>to</strong> make contact with the surface of the opposing gill.The dashed<br />

arrow indicates the trajec<strong>to</strong>ry (sporabola) taken by a projected spore.


STRUCTURE AND MORPHOGENESIS OF BASIDIOCARPS<br />

523<br />

Fig19.8 Modes of gill attachment <strong>to</strong> the stipe, and<br />

their terminology.<br />

of gill maturation begins at the cap margin and<br />

passes slowly upwards and inwards. After the<br />

basidia at the lower edge of the gill have discharged<br />

their spores, the gill tissue undergoes<br />

au<strong>to</strong>lysis (deliquescence) in<strong>to</strong> an inky black<br />

liquid which drips away from the cap. The gravitropic<br />

gill curvature characteristic of aequihymenial<br />

types is absent, but the stipe may still<br />

curve <strong>to</strong> bring the gills in<strong>to</strong> an approximately<br />

vertical position.<br />

An important criterion for identification of<br />

gill-bearing agarics in the field is the way in<br />

which the lamellae are attached <strong>to</strong> the stipe<br />

(Fig. 19.8). Gills are said <strong>to</strong> be free if their blade<br />

does not <strong>to</strong>uch the stipe. Adnate gills show<br />

attachment <strong>to</strong> the stipe with their entire base,<br />

whereas adnexed gills are attached only partially<br />

<strong>to</strong> the stipe. In sinuate gills, the gill margin<br />

shows an S-shaped curve near the point of<br />

junction with the stipe. Decurrent gills run<br />

down the surface of the stipe.<br />

19.2.4 The hymenophoral trama<br />

The hymenophore structures of some<br />

Homobasidiomycetes are shown in Fig. 19.9. In<br />

most hymenophores, there is a central group of<br />

hyphae running from the underside of the cap <strong>to</strong><br />

the tip of the gill, pore or spine, and these<br />

hyphae are collectively called the hymenophoral<br />

trama. Various distinctive types have been<br />

recognized by Reijnders and Stalpers (1992), but<br />

these do not correlate well with the phylogeny of<br />

the Homobasidiomycetes as shown in Fig. 19.2.<br />

1. In the trame<strong>to</strong>id type (e.g. in Lentinus,<br />

Fistulina and Schizophyllum), development begins<br />

with a bundle of parallel hyphae from which<br />

branches grow out in various directions and may<br />

become interwoven.<br />

2. In the cantharelloid type, the developing<br />

hymenophore is at first smooth and covered by a<br />

palisade of hyphae which will form<br />

the hymenium. Later, locally increased<br />

activity in the subhymenium may result<br />

in irregular ridges (Serpula), more regular veins<br />

(Cantharellus, Craterellus) or gills (Hygrophoropsis).<br />

3. The bole<strong>to</strong>id type is shown by Boletus and<br />

relatives (Fig. 19.9a). The first stage is made up of<br />

divergent hyphae which form a narrow central<br />

layer (mediostratum), followed by swelling and<br />

differentiation of the cells <strong>to</strong> form a lateral<br />

stratum. The diverging hyphae of the lateral<br />

strata curve sharply outwards <strong>to</strong> form the<br />

hymenium, and the term divergent trama is<br />

therefore used <strong>to</strong> describe this type of hymenophore.<br />

A subhymenium made up of narrow<br />

interwoven hyphae is also present.<br />

4. The agaricoid type is by far the most<br />

common and includes the coprinoid, russuloid,<br />

agaricoid, pluteoid and amani<strong>to</strong>id subtypes. In<br />

cross-section, gills usually appear differentiated<br />

in<strong>to</strong> a central trama, a subhymenium and the<br />

hymenium (Fig. 19.9b). The hyphae in the trama<br />

often run parallel <strong>to</strong> each other, but in the<br />

pluteoid subtype (e.g. Volvariella and Pluteus), the<br />

wide hyphae making up the trama are arranged<br />

in a V-shaped pattern (Fig. 19.9d), and this is<br />

called the inverted trama. In the russuloid<br />

subtype (Russula and Lactarius), the mature<br />

trama is said <strong>to</strong> be intermixed because in<br />

addition <strong>to</strong> the narrow generative hyphae it<br />

contains swollen cells (sphaerocysts) of very<br />

different width (Fig. 19.9c). The amani<strong>to</strong>id type<br />

shows a central mediostratum made up of<br />

narrow generative hyphae from which wider<br />

subhymenial elements diverge (Fig. 19.9e). This<br />

type of trama is described as bilateral.


524 HOMOBASIDIOMYCETES<br />

Fig19.9 Different types of hymenophoral trama. (a) The bole<strong>to</strong>id type shown by Boletus edulis. L.S. immature pore showing the<br />

divergent trama. (b e) Vertical sections of gills of various agarics of the aequi-hymenial type showing different types of<br />

hymenophore. (b) The agaricoid type shown in Flammulina velutipes. (c) The russuloid subtype shown by Russula cyanoxantha.Note<br />

the globose sphaerocysts. (d) The pluteoid subtype shown by Pluteus cervinus. Note the converging (V-shaped) arrangement of the<br />

cells making up the trama (inverted trama) and the hooked cystidia. (e) The amani<strong>to</strong>id subtype.Vertical section of gill of Amanita<br />

rubescens, showing the bilateral hymenophoral trama. (c) and (d) <strong>to</strong> same scale.


IMPORTANCE OF HOMOBASIDIOMYCETES<br />

525<br />

19.2.5 The hymenium<br />

The ends of the tramal hyphae turn outwards<br />

<strong>to</strong> form a distinct layer of shorter cells, the<br />

subhymenium, which lies immediately beneath<br />

the hymenium of a palisade-like layer of ripe<br />

basidia, developing basidia (basidioles) and sometimes<br />

other structures such as cystidioles,<br />

cystidia and paraphyses. All these represent the<br />

terminal cells of hyphae making up the fruit<br />

body and are therefore homologous with basidia.<br />

Cystidioles are thin-walled, sterile elements of<br />

the hymenium, about the same diameter as the<br />

basidia, and usually protruding only slightly<br />

from the hymenial surface.<br />

Cystidia are more varied. They are often<br />

enlarged conical or cylindrical cells which may<br />

arise in the hymenium along with the basidia<br />

(hymenial cystidia) or sometimes deeper, for<br />

example in the trama (tramal cystidia). Fusion<br />

of paired nuclei within cystidia has been<br />

reported and in some cases 2 4 basidiospores<br />

have been observed <strong>to</strong> develop on cystidia. In<br />

Coprinus and its allies, the cystidia on the face of<br />

the gill may stretch across the space between the<br />

gills (Fig. 19.7). The tip of a cystidium makes<br />

contact with a cell on the opposing hymenium,<br />

and this cell, termed a cystesium (from cystidium<br />

þ Lat. haerare ¼ <strong>to</strong> adhere), then differentiates<br />

by developing a granular, vacuolated<br />

cy<strong>to</strong>plasm and becomes cemented <strong>to</strong> the tip of<br />

the cystidium, so that the cystidium bridging the<br />

gill cavity becomes firmly attached <strong>to</strong> both<br />

hymenia (Horner & Moore, 1987; Moore, 1998).<br />

The role of the hymenial cystidia in Coprinus<br />

was initially thought <strong>to</strong> be that of buttresses<br />

helping <strong>to</strong> space the inaequi-hymenial gills<br />

apart. However, it is now believed that the<br />

cystidium cystesium pairs function as tension<br />

elements, balancing the stretching forces generated<br />

by expansion growth of the cap and helping<br />

<strong>to</strong> straighten out the thin lamellae which are<br />

folded during early development (Chiu & Moore,<br />

1990a). In Pluteus the tramal cystidia bear hooklike<br />

tips (Fig. 19.9d) whose function is not<br />

unders<strong>to</strong>od. The suggestion that they might<br />

deter animals such as slugs from eating the gill<br />

tissue is not supported by feeding experiments<br />

(Buller, 1924). In Volvariella bombycina, cystidia on<br />

the face of the gill have aqueous drops adhering<br />

<strong>to</strong> them and are thus believed <strong>to</strong> be secre<strong>to</strong>ry in<br />

function, releasing water vapour in<strong>to</strong> the space<br />

surrounding the basidia (Chiu & Moore, 1990b).<br />

Various terms have been used <strong>to</strong> describe cystidia<br />

from different parts of the basidiocarp. Those on<br />

the gill face are termed facial cystidia or<br />

pleurocystidia, those on the gill margin cheilocystidia.<br />

Cystidia are not confined <strong>to</strong> the hymenium.<br />

Similar structures have been found on the<br />

surface of the pileus (pileocystidia) and the stipe<br />

(caulocystidia). We are still ignorant of the<br />

function of many of these structures. For a<br />

fuller discussion see Singer (1986) and<br />

Clémençon (2004).<br />

Paraphyses are present in the hymenium of<br />

Coprinus. They develop on branches from the<br />

same hyphae which produce basidia (Fig. 19.7),<br />

arising after the young basidia are committed <strong>to</strong><br />

meiosis. The tips of the paraphyses force themselves<br />

in<strong>to</strong> the hymenium between developing<br />

basidia, grow <strong>to</strong> about two-thirds of the height of<br />

a basidium, become spherical and enlarge, thus<br />

playing an important role in the expansion of<br />

the maturing basidiocarp (Rosin & Moore, 1985;<br />

Moore, 1998).<br />

19.3 Importance of<br />

homobasidiomycetes<br />

Homobasidiomycetes have a significant impact<br />

on our lives. The fruit bodies of thousands<br />

of species are potentially edible, although<br />

only some 40 species have been cultivated.<br />

According <strong>to</strong> worldwide production data for<br />

the year 1997 (Chang & Miles, 2004), the most<br />

popular cultivated mushrooms are Agaricus<br />

bisporus (2 000 000 t per annum), Lentinula edodes<br />

(1 500 000 t), Pleurotus spp. (875 600 t), Flammulina<br />

velutipes (284 000 t) and Volvariella volvacea<br />

(180 000 t). Mushroom production is expanding<br />

rapidly, especially in the Far East. However,<br />

some of the most prized edible mushrooms are<br />

mycorrhizal associates of trees which cannot yet<br />

be cultivated away from their host. Mycorrhizal<br />

associations involving Homobasidiomycetes are<br />

prominent in forest situations and are dealt<br />

with in more detail below and also on p. 581.


526 HOMOBASIDIOMYCETES<br />

An equally important ecological role played by<br />

the Homobasidiomycetes in the global environment<br />

is that of saprotrophs involved in the<br />

decomposition of the two most abundant organic<br />

carbon sources, cellulose and lignin, thereby<br />

releasing nutrients locked up in wood and leaf<br />

litter. The degradation of wood is achieved in<br />

two different ways, white-rot and brown-rot,<br />

and these are described briefly on pp. 527 532.<br />

A few wood-rotting species are plant pathogens,<br />

e.g. Armillaria mellea, Phellinus noxius and Crinipellis<br />

perniciosa, whereas others, notably the dry rot<br />

fungus Serpula lacrymans, cause economic damage<br />

of timber built in<strong>to</strong> houses.<br />

19.3.1 Ec<strong>to</strong>mycorrhiza<br />

Many members of the Homobasidiomycetes<br />

(including gasteromycetes), as well as a few<br />

Ascomycota such as the truffles (Tuber spp.; see<br />

p. 423), form ec<strong>to</strong>mycorrhizal associations with<br />

coniferous and broad-leaved trees. These are<br />

distinguished from the vesicular arbuscular<br />

mycorrhiza between herbaceous plants and<br />

Zygomycota (see p. 217) in several key features<br />

which have been reviewed by Smith and Read<br />

(1997) and Peterson et al. (2004). The most<br />

immediately obvious is that the bulk of the<br />

fungal biomass is located outside the plant root,<br />

hence the term ec<strong>to</strong>mycorrhiza. The colonization<br />

of a tree root by an ec<strong>to</strong>mycorrhizal fungus has<br />

reciprocal morphogenetic effects. Lateral roots<br />

show stunted growth accompanied by increased<br />

branching which is often dicho<strong>to</strong>mous<br />

(Fig. 19.10a) in response <strong>to</strong> fungal colonization,<br />

and they are covered by a thick sheath of hyphae,<br />

the mantle (Fig. 19.10b). There is also a limited<br />

colonization of the root cortex in the shape of<br />

the Hartig net, a system of unusually richly<br />

branched intercellular hyphae (Fig. 19.10b). The<br />

sequence of colonization events probably starts<br />

with hyphae being initially attracted <strong>to</strong> root tips<br />

by their exudates or possibly those of microorganisms<br />

associated with the rhizosphere, such<br />

as fluorescent pseudomonads (Garbaye, 1994;<br />

Smith & Read, 1997). Following contact with a<br />

root hair, hyphae grow alongside it until they<br />

meet the surface of the main root (Thomson<br />

et al., 1989). There, morphogenetic changes are<br />

initiated, such as hyphal branching and anas<strong>to</strong>mosis<br />

which lead <strong>to</strong> the establishment of the<br />

mantle (Massicotte et al., 1987). These changes<br />

may result from the specific recognition of wall<br />

surface molecules between the root and fungus<br />

hypha (Giollant et al., 1993; Lapeyrie & Mendgen,<br />

1993). Outside the mantle, the mycelium<br />

may extend in<strong>to</strong> the soil by a few centimetres,<br />

or much further if the fungus is capable of<br />

forming mycelial cords (see p. 581). The roots of<br />

different plants in complex forest ecosystems<br />

may be linked by common ec<strong>to</strong>mycorrhizal<br />

fungi, and there may be a net transfer of<br />

carbon from sunlit plants <strong>to</strong> those growing in<br />

the shade (Leake et al., 2004). Mineral nutrients,<br />

notably phosphate, as well as water are transported<br />

from the fungus <strong>to</strong> the plant.<br />

Carbohydrates travel the opposite way. Sucrose,<br />

the main transport carbohydrate in plants,<br />

is secreted in<strong>to</strong> the apoplast and hydrolysed <strong>to</strong><br />

give fruc<strong>to</strong>se and glucose. The latter is taken up<br />

by the fungus and converted <strong>to</strong> glycogen, trehalose<br />

or polyols (Smith & Read, 1997). Some 10% of<br />

the net pho<strong>to</strong>synthetic assimilate may be allocated<br />

<strong>to</strong> mycorrhizal fungi which make up<br />

20 30% of the microbial biomass in forest soils<br />

(Leake et al., 2004). Fruit bodies are a major sink<br />

for translocated carbon. Colourless (achlorophyllous)<br />

plants may obtain their carbon by plugging<br />

in<strong>to</strong> the mycorrhizal network, a strategy known<br />

as mycoheterotrophy (Smith & Read, 1997).<br />

Fossil records of ec<strong>to</strong>mycorrhizal associations<br />

date back some 50 million years (LePage et al.,<br />

1997), although they are more likely <strong>to</strong> be<br />

around 200 million years old (Cairney, 2000).<br />

Mycorrhizae of this kind are particularly prominent<br />

in nutrient-poor or dry soils. Many ec<strong>to</strong>mycorrhizal<br />

fungi have retained the capacity<br />

<strong>to</strong> produce hydrolytic enzymes and are capable<br />

of solubilizing, for example, phosphorus and<br />

nitrogen from complex sources (Perez-Moreno &<br />

Read, 2000). They are therefore often associated<br />

with humus layers in which saprotrophic<br />

species also grow. Ec<strong>to</strong>mycorrhizal fungi may<br />

show a greater or lesser degree of host specificity;<br />

for instance, Suillus grevillei is associated almost<br />

exclusively with larch (Larix spp.), whereas<br />

Amanita muscaria (fly agaric), Boletus edulis<br />

(cep or penny bun) or Cantharellus cibarius


IMPORTANCE OF HOMOBASIDIOMYCETES<br />

527<br />

Fig19.10 Ec<strong>to</strong>mycorrhiza on beech roots collected from humus. (a) General view. Note the stunted growth and dicho<strong>to</strong>mous<br />

branching of the lateral roots which are entirely covered by the mantle. (b) Transverse section through such a mycorrhizal root tip.<br />

A mantle several hyphae thick has formed a sheath around the epidermis. Branching hyphae grow between the outer cortical cells<br />

(arrows) <strong>to</strong> form the Hartig net. (a) kindly provided by A.E. Ashford.<br />

(chanterelle) are found under both broad-leaved<br />

and coniferous trees.<br />

The latter two species <strong>to</strong>gether with<br />

Tricholoma matsutake are the three most valuable<br />

mycorrhizal Homobasidiomycetes in commercial<br />

terms, with a combined annual crop value of<br />

over US$ 2 billion (Hall et al., 2003). Considerable<br />

efforts have been made <strong>to</strong> develop methods for<br />

their cultivation, but success has been limited.<br />

Whilst it is possible <strong>to</strong> cultivate mycelium of<br />

these species in vitro on agar media and <strong>to</strong><br />

inoculate seedlings of host trees, the crops of<br />

fruit bodies after outplanting have been disappointing<br />

or al<strong>to</strong>gether absent (Wang & Hall,<br />

2004). The situation is more encouraging for the<br />

ec<strong>to</strong>mycorrhizal ascomycete truffles Tuber melanosporum<br />

and T. magnatum (see p. 423), but even<br />

with these species the establishment of commercial<br />

plantations has failed <strong>to</strong> reverse the decline<br />

in overall harvests due <strong>to</strong> overcollection and<br />

destruction of natural habitats.<br />

Fortunately, there are several ec<strong>to</strong>mycorrhizal<br />

species which, although inedible, can be<br />

used as reliable inoculum <strong>to</strong> promote the<br />

growth of trees in challenging environmental<br />

situations. Probably the most successful group<br />

is the gasteromycete genus Pisolothus, which is<br />

used extensively in afforestation programmes<br />

(see p. 581).<br />

19.3.2 Degradation of wood<br />

The complex architecture of wood is summarized<br />

in Fig. 19.11. The middle lamella separating<br />

adjacent cells consists of a-(1,4)-linked galacturonic<br />

acid polymers collectively called pectins, in<br />

which the carboxylic acid group may be derivatized<br />

by methylation or calcium salt formation.<br />

The primary wall consists chiefly of cellulose<br />

made up of b-(1,4)-linked glucose units.<br />

Hemicelluloses are also present; these comprise<br />

several heterogeneous polymers with a backbone<br />

of glucose, mannose or xylose in b-(1,4)-linkage,


528 HOMOBASIDIOMYCETES<br />

Fig19.11 Carbon-containing polymers and their localization in woody tissue.Lignin consists of a complex polymer (modified from Adler,1977) of the three phenylpropanoid units shown,<br />

which may be linked in numerous different combinations.


IMPORTANCE OF HOMOBASIDIOMYCETES<br />

529<br />

with various sugars or uronic acids added as side<br />

chains. Since pectin, cellulose and hemicelluloses<br />

are all polymerized by enzymes in a regular<br />

fashion, they can also be degraded by hydrolytic<br />

Fig19.12 Wood rot symp<strong>to</strong>ms. (a) White-rot of beech caused<br />

byTrametes hirsuta (left). Note the bleached appearance of the<br />

branch interior at the broken surface. Beech wood attacked<br />

by a brown-rot is also shown for comparison (right). (b) Trunk<br />

segments of Picea abies attacked by Fomi<strong>to</strong>psis pinicola.The<br />

brown-rot had caused a hollowing of the standing tree prior<br />

<strong>to</strong> felling. (c) Log of Picea abies colonized by Fomi<strong>to</strong>psis pinicola,<br />

showing brown rot symp<strong>to</strong>ms.The wood has cracked in<strong>to</strong><br />

cube-like fragments which are easily ground in<strong>to</strong> a powder.<br />

enzymes, although accessibility problems may<br />

prevent this from happening in situ (see below).<br />

The architecture of secondary plant walls is<br />

radically different. Apart from hemicelluloses<br />

as a minor component, the principal building<br />

material is lignin, a polymer of aromatic<br />

alcohols which are cross-linked in a random<br />

fashion by free radical reactions. Lignin therefore<br />

has a complex, non-repetitive three-dimensional<br />

structure resistant <strong>to</strong> direct attack by<br />

enzymes. Only few basidiomycetes are able <strong>to</strong><br />

mineralize lignin <strong>to</strong> H 2 O and CO 2 , and this is<br />

achieved by oxidative rather than hydrolytic<br />

enzymes. The result is that wood undergoes<br />

white-rot (see Plate 10a; Fig. 19.12a), i.e. it<br />

appears bleached because all its components<br />

are degraded more or less simultaneously, or<br />

lignin degradation precedes attack on cellulose.<br />

In contrast, many fungi which degrade cellulose<br />

leave behind the lignin component of wood<br />

which turns brown upon oxidation, and this type<br />

of decay is therefore called brown-rot. The<br />

removal of cellulose destroys the structural<br />

integrity of wood so that the lignin cracks in<strong>to</strong><br />

cubes (Fig. 19.12c) and ultimately breaks up in<strong>to</strong><br />

a powder which becomes incorporated in<strong>to</strong><br />

humus. Living trees may survive infection by<br />

brown-rot fungi if this is confined <strong>to</strong> the heartwood<br />

in the core of the trunk, leaving a<br />

sufficiently strong cylinder of sound wood<br />

(Fig. 19.12b). Prolonged or repeated attacks by<br />

such wood-rotting basidiomycetes are responsible<br />

for the hollowing of trunks observed in<br />

many his<strong>to</strong>ric trees.<br />

Brown-rot<br />

The hyphae of brown-rot fungi colonizing wood<br />

through its lignin-encased tubular cavities face<br />

the major problem of gaining access <strong>to</strong> degradable<br />

substrates. One point of attack is the middle<br />

lamella, which is exposed by simple or bordered<br />

pits in the secondary cell walls of adjacent cells,<br />

and polygalacturonases are indeed produced<br />

by many brown-rot fungi (Green & Clausen,<br />

1999, 2003). A typical feature of brown-rots is<br />

the production of oxalic acid, which may be<br />

important in chelating calcium ions released at<br />

potentially <strong>to</strong>xic concentrations by the degradation<br />

of calcium pectate. Calcium oxalate crystals


530 HOMOBASIDIOMYCETES<br />

are common on the surfaces of many fungal<br />

hyphae. Oxalic acid may also chelate other ions<br />

such as Cu 2þ , which is used as a preservative for<br />

the treatment of wood. The oxalic acid concentration<br />

may bring about strongly acidic conditions,<br />

sometimes as low as pH 1.7, which would<br />

be sufficient for acid hydrolysis of pectin and<br />

even cellulose (Green et al., 1991). Both endoand<br />

exo-enzymes of the cellulase complex are<br />

generally produced by brown-rot fungi (Hegarty<br />

et al., 1987), and these may act in the usual<br />

synergistic way <strong>to</strong> convert cellulose <strong>to</strong> oligosaccharides<br />

and thence <strong>to</strong> glucose (Radford et al.,<br />

1996). Hemicellulases are thought <strong>to</strong> act in a<br />

similar way. However, these enzymes may<br />

not gain access <strong>to</strong> their substrate where it is<br />

masked by the lignin-containing secondary wall.<br />

Although the detailed mechanism of brown-rot<br />

decay is still unknown, there is evidence that<br />

at least the initial attack on cellulose<br />

(and hemicellulose) is mediated by small molecules<br />

capable of penetrating the lignin layer.<br />

These may be the hydronium ion (H 3 O þ ) generated<br />

by oxalic acid in water, or another molecule<br />

such as the hydroxyl radical (HO) released from<br />

hydrogen peroxide (H 2 O 2 ) by the Fen<strong>to</strong>n reaction<br />

(Fe 2þ þ H 2 O 2 ! Fe 3þ þ HO þ HO ). The<br />

mechanism of brown-rot decay is all the more<br />

mysterious because many brown-rot fungi do,<br />

in fact, produce enzymes capable of degrading<br />

lignin (Mtui & Nakamura, 2004), but these may<br />

have other functions, such as the de<strong>to</strong>xification<br />

of antimicrobial phenolics which are often<br />

present in wood at high concentrations<br />

(Rabinovich et al., 2004).<br />

White-rot<br />

The suite of enzymes required <strong>to</strong> break down<br />

lignin has been most thoroughly examined in<br />

the white-rot fungus Phanerochaete chrysosporium<br />

(see de Jong et al., 1994a; Heinzkill & Messner,<br />

1997). There are numerous conflicting ideas<br />

about the enzymology of lignin degradation,<br />

and we can only generalize here. Initial attack on<br />

lignin is mediated by lignin peroxidases (LiP)<br />

and/or manganese peroxidases (MnP) which do<br />

not themselves enter the lignin layer. Instead,<br />

small diffusible molecules probably act as redox<br />

charge carriers between the enzymes and their<br />

substrate. A possible reaction scheme for LiP is<br />

shown in Fig. 19.13 (Heinzkill & Messner, 1997;<br />

ten Have & Teunissen, 2001). The enzyme consists<br />

of a single polypeptide chain and a pro<strong>to</strong>porphyrin<br />

IX (haem) group which is buried deep<br />

inside the enzyme, accessible <strong>to</strong> small diffusible<br />

molecules through a narrow pore. LiP loses two<br />

electrons when it reduces H 2 O 2 <strong>to</strong> water. This<br />

highly oxidized LiP I state is returned <strong>to</strong> the<br />

ground state in two steps, each associated with<br />

the one-electron oxidation of a reductant in<strong>to</strong><br />

its cation radical. Veratryl alcohol, produced<br />

in abundance by most white-rot fungi, is such<br />

a reductant. The two cation radicals leave the<br />

enzyme and either themselves attack the lignin<br />

structure, or pass on their charge <strong>to</strong> other small<br />

molecules. Either way, the extraction of one<br />

electron from an aromatic ring of lignin generates<br />

a structure reacting both as a cation and as a<br />

radical, leading <strong>to</strong> numerous possible degradation<br />

products. The conversion of LiP in<strong>to</strong> LiP I<br />

requires H 2 O 2 which is generated by extracellular<br />

enzymes such as glucose oxidase or aryl<br />

alcohol oxidase. The latter uses organohalogens<br />

such as 3-chloroanisylalcohol as substrates, and<br />

these may accumulate in the environment<br />

colonized by white-rot fungi (de Jong et al.,<br />

1994b), even though organohalogens have traditionally<br />

been regarded as man-made (anthropogenic)<br />

environmental pollutants.<br />

Manganese peroxidase (MnP) is closely related<br />

<strong>to</strong> LiP in its protein structure and its catalytic<br />

cycle, except that the co-fac<strong>to</strong>r is Mn 2þ instead of<br />

veratryl alcohol. The oxidized Mn 3þ may be<br />

stabilized by organic acids en route <strong>to</strong> the<br />

lignin substrate, where it catalyses a one-electron<br />

oxidation either directly or via charge transfer<br />

molecules, followed by recycling of Mn 2þ back <strong>to</strong><br />

the MnP enzyme. Copper-containing laccases are<br />

a third group of enzymes attacking lignin, and<br />

these are produced by a range of fungi much<br />

wider than that causing white-rot. By reducing<br />

O 2 <strong>to</strong> H 2 O, laccases are capable of performing a<br />

four-electron oxidation either of a redox carrier<br />

or the final substrate itself. In this way, laccases<br />

may be able <strong>to</strong> degrade smaller lignin fragments,<br />

although they are probably incapable of a direct<br />

attack on intact lignin, due <strong>to</strong> steric problems.<br />

The precise role of laccases in lignin degradation


IMPORTANCE OF HOMOBASIDIOMYCETES<br />

531<br />

Fig19.13 Reaction cycle of lignin peroxidase.The reduction of H 2<br />

O 2<br />

<strong>to</strong> H 2<br />

O withdraws two electrons from the ground-state<br />

enzyme (LiP), one from the ferric ion <strong>to</strong> give the ferryl ion (Fe4þ), and the other from the pro<strong>to</strong>porphyrin group itself which is<br />

converted <strong>to</strong> the porphyrin cation radical.This results in the highly oxidized LiP I state which catalyses two separate one-electron<br />

oxidations of a reductant, e.g. veratryl alcohol (VA), becoming reduced via the LiP II state <strong>to</strong> the ground state.TheVA cation<br />

radical is regenerated when it withdraws one electron from lignin itself or from other molecules serving as redox charge carriers<br />

(not shown). H 2<br />

O 2<br />

can be generated by several means, including the oxidation of 3-chloro-anisylalcohol (CA alc) <strong>to</strong> give its<br />

corresponding aldehyde (CA ald).<br />

has not yet been fully characterized, although it<br />

is likely <strong>to</strong> be pivotal as most, if not all, white-rot<br />

fungi produce them, and in some white-rot<br />

species such as Pycnoporus cinnabarinus they may<br />

be the only kind of ligninolytic enzyme present<br />

(Eggert et al., 1997). All three lignin-attacking<br />

enzymes are usually secreted as multiple<br />

isoforms which differ in their pH optima or<br />

catalytic properties (Kirk & Cullen, 1998). It<br />

should be noted that lignin degradation is<br />

always co-metabolic, i.e. white-rot fungi cannot<br />

utilize lignin as the sole source of carbon or<br />

energy, and they probably remove it for the<br />

purpose of gaining access <strong>to</strong> more easily<br />

degraded substrates such as cellulose. Although<br />

white-rot fungi are generally considered <strong>to</strong> be<br />

more ‘efficient’ than brown-rots, the latter, in<br />

circumventing the effort of lignin degradation,<br />

probably obtain more glucose per unit metabolic<br />

effort.<br />

The attack on lignin by means of charge<br />

transfer molecules being a non-specific reaction<br />

mechanism, ligninolytic enzymes will also<br />

degrade other substances, including man-made<br />

recalcitrant environmental pollutants such<br />

as polycyclic aromatic hydrocarbons, organohalogens,<br />

or the explosive TNT (see Gadd, 2001).<br />

White-rot fungi are therefore being investigated<br />

extensively for their potential in the bioremediation<br />

of contaminated environments. Another


532 HOMOBASIDIOMYCETES<br />

biotechnological application is the use of oxidative<br />

enzymes for the bleaching of wood pulp for<br />

paper production, or of textile dyes.<br />

There is a further twist <strong>to</strong> the tale. Many<br />

fungal metabolites, including veratryl alcohol<br />

and 3-chloro-anisylalcohol (see Fig. 19.13), contain<br />

methoxy groups or methyl esters, whose<br />

methyl groups are added late during biosynthesis.<br />

Various molecules may act as donors<br />

of methyl groups, including chloromethane<br />

(CH 3 Cl), a known greenhouse gas. A few woodrotting<br />

fungi, notably Phellinus spp., have<br />

gained no<strong>to</strong>riety because they produce vastly<br />

more chloromethane than they require for<br />

their biosynthetic machinery. Watling and<br />

Harper (1998) estimated that atmospheric chloromethane,<br />

which is mainly of natural origin,<br />

accounts for 15 20% of the chlorine-mediated<br />

ozone destruction in the stra<strong>to</strong>sphere, and that<br />

86% of the <strong>to</strong>tal fungal chloromethane contribution<br />

can be attributed <strong>to</strong> the genus Phellinus.<br />

19.4 Euagarics clade<br />

As shown in Table 19.1 and Fig. 19.2, the<br />

euagarics clade includes not only forms with<br />

lamellate (i.e. agaricoid) basidiocarps formerly<br />

classified in the Agaricales, but also forms with<br />

other hymenial configurations. This means that<br />

there are few, if any, reliable gross morphological<br />

characters by which members of the euagarics<br />

clade can be recognized. It is probable that<br />

agaricoid basidiocarp types have evolved repeatedly<br />

(Hibbett et al., 1997b). Kirk et al. (2001) have<br />

recognized 26 families, 347 genera and about<br />

10 000 species, but cautioned that there are<br />

difficulties in defining these families at present.<br />

In the account of some common genera which<br />

follows, the estimated numbers of species have<br />

been taken from Kirk et al. (2001) unless otherwise<br />

indicated.<br />

19.4.1 Agaricaceae<br />

The Agaricaceae are one of the most diverse<br />

families of the Agaricales, estimated <strong>to</strong> contain<br />

over 50 genera and some 900 species. The spore<br />

print (i.e. the accumulation of spores projected<br />

from a basidiocarp) may be white or coloured.<br />

There are also variations in the structure of the<br />

hymenophoral trama and the surface of the<br />

pileus. Partial and/or universal veils are usually<br />

present. Despite these variations, molecular<br />

evidence supports the view that the core genera<br />

of this family, including Agaricus and Lepiota, are<br />

monophyletic (Moncalvo et al., 2000, 2002;<br />

Vellinga, 2004).<br />

Agaricus (c. 200 spp.)<br />

This is a large genus of fungi often synonymized<br />

with the term ‘mushrooms’, distributed mainly<br />

in temperate regions. Agaricus spp. are saprotrophs,<br />

growing in pastures and woodland litter.<br />

The mycelium is perennial and some species, e.g.<br />

A. arvensis (horse mushroom), A. xanthodermus<br />

(yellow stainer) and A. tabularis may fruit in rings.<br />

From measurements of the annual rate of<br />

increase in diameter, a ring of A. tabularis 60 m<br />

in diameter was estimated <strong>to</strong> be 250 years old.<br />

Agaricus basidiocarps are moderate <strong>to</strong> large in<br />

size, generally firm and white but sometimes<br />

changing colour upon bruising (e.g. in A. xanthodermus).<br />

There is a ring on the stem (two rings in<br />

A. bi<strong>to</strong>rquis), but no volva. The gills are at first<br />

pink due <strong>to</strong> the colour of the cy<strong>to</strong>plasm of the<br />

spores, but, as the spores mature, their walls<br />

darken <strong>to</strong> a purple-brown, the colour of the spore<br />

print. Many species have edible basidiocarps<br />

and some are prized as food, e.g. A. campestris<br />

(field mushroom), A. arvensis, A. macrosporus and<br />

A. silvaticus. Agaricus bisporus (Fig. 19.14a), now<br />

sometimes called A. brunnescens, is the cultivated<br />

white but<strong>to</strong>n mushroom (see below) and is<br />

occasionally found in nature on manure heaps,<br />

garden waste and roadsides, and under Cupressus<br />

in coastal areas of California and France. The<br />

basidiocarps of A. xanthodermus may cause gastrointestinal<br />

upsets in some people, probably due<br />

<strong>to</strong> the presence of phenol which also gives the<br />

fruit bodies an unpleasant smell of carbolic acid<br />

(Gill & Strauch, 1984).<br />

Cultivation of Agaricus bisporus<br />

The white but<strong>to</strong>n mushroom has been cultivated<br />

for its edible fruit bodies for almost four<br />

centuries since collec<strong>to</strong>rs discovered that its<br />

spawn could be used <strong>to</strong> inoculate compost. The


EUAGARICS CLADE<br />

533<br />

Fig19.14 Basidiocarps in the euagarics clade (1). (a)The commercial mushroom, Agaricus bisporus. (b) The parasol mushroom,<br />

Macrolepiota procera. In mature specimens the ring is moveable. (c) Coprinus comatus, the shaggy inkcap or lawyer’s wig. (d) Coprinus<br />

cinereus fruiting in the labora<strong>to</strong>ry on a dung/straw mixture. (e) Pleurotus ostreatus,theoystermushroom.<br />

widespread cultivars used in the mushroom<br />

industry originated in France, where it is called<br />

champignon de Paris or champignon de couche.<br />

There are probably about seven such ancestral<br />

lineages (Callac, 1995). Specimens growing in<br />

nature belong <strong>to</strong> two distinct groups, namely<br />

genetically diverse indigenous populations and<br />

more uniform ‘cultivar-like’ populations<br />

which probably represent escapes from cultivation<br />

(Xu et al., 1997). Mushroom cultivation is<br />

now an important industry in temperate countries,<br />

and technical details have been described<br />

by van Griensven (1988) and Chang and Miles<br />

(2004). In 1996 global production exceeded


534 HOMOBASIDIOMYCETES<br />

2 million <strong>to</strong>nnes (Moore & Chiu, 2001). In areas<br />

with a warmer climate, e.g. in Southern Europe,<br />

A. bi<strong>to</strong>rquis is also grown.<br />

Mushrooms were originally cultivated in<br />

caves, but <strong>to</strong>day most are grown in specially<br />

constructed sheds. The basic substratum is a<br />

composted straw/manure mixture (Straatsma<br />

et al., 1995). The heap of compost is kept outdoors<br />

where it heats up naturally <strong>to</strong> 80°C over a period<br />

of 2 weeks, with occasional turning. The heat<br />

produced during composting selects for the<br />

growth of specialized thermophilic microbes,<br />

the most important of which for mushroom<br />

production is the anamorphic mould Scytalidium<br />

thermophilum (syn. Humicola insolens) (Straatsma &<br />

Samson, 1993). The yield of mushrooms from<br />

a compost inoculated with S. thermophilum is<br />

doubled as compared with pasteurized controls<br />

(Straatsma et al., 1994a,b) because the A. bisporus<br />

mycelium can grow partly on the residues<br />

associated with the activities of S. thermophilum<br />

and partly on its living mycelium and conidia<br />

(Bilay & Lelley, 1997). The heat produced during<br />

composting destroys many competing microbes,<br />

and further destruction is achieved by a pasteurization<br />

process in which the compost, now<br />

moved in<strong>to</strong> a shed, is heated by steam <strong>to</strong> a<br />

temperature of 60°C for 8 h which minimizes<br />

diseases caused later by other fungi.<br />

After cooling, the compost is inoculated with<br />

mushroom spawn pre-grown on sterilized cereal<br />

grains. The inoculated compost is incubated at<br />

25°C for 2 weeks, and fruiting is induced by<br />

covering the colonized spawn with a 3 5cm<br />

deep layer of casing soil or a special mixture of<br />

moist peat and chalk. The mycelium grows up<br />

through the casing layer and forms anas<strong>to</strong>mosing<br />

mycelial cords from which ‘pinheads’<br />

develop at the surface, some of which expand<br />

<strong>to</strong> form mature basidiocarps. In addition <strong>to</strong> any<br />

physical or chemical effect associated with the<br />

casing layer, there is also a biological component.<br />

This is shown by the fact that sterilized<br />

casing soil is far less effective than natural soil in<br />

encouraging fruiting. Re-inoculation of bacteria<br />

isolated from casing soil in<strong>to</strong> sterile casing has<br />

shown that Pseudomonas spp. and especially<br />

P. putida are effective in inducing fruiting<br />

(Hayes et al., 1969). The growth of P. putida,<br />

in turn, is selectively stimulated by volatile<br />

substances such as ethanol emanating from<br />

the mushroom compost. Bacteria accumulate<br />

around Agaricus hyphae in a zone sometimes<br />

termed the hyphosphere.<br />

Bacteria which are active in inducing basidiocarp<br />

formation might achieve their effect<br />

in two ways. Possibly, they induce starvation<br />

of A. bisporus hyphae in a manner similar <strong>to</strong><br />

fungistasis in soil, where the depletion of<br />

nutrients by competing microbes inhibits<br />

fungal spore germination. An alternative idea<br />

is that bacteria in the casing layer<br />

absorb substances which retard the fruiting of<br />

A. bisporus. This idea is supported by observations<br />

that fruiting can be induced by replacing the<br />

casing layer with activated charcoal, believed <strong>to</strong><br />

adsorb inhibi<strong>to</strong>ry substances (Long & Jacobs,<br />

1974; Flegg & Wood, 1985; Noble et al., 2003).<br />

The isolation of putative inhibi<strong>to</strong>rs is a difficult<br />

task, but one substance, 1-octen-3-ol, has been<br />

shown <strong>to</strong> inhibit primordium formation in plate<br />

culture.<br />

Two weeks after application of the casing<br />

layer, fruiting is stimulated by forced<br />

air ventilation which lowers the temperature<br />

<strong>to</strong> 20°C and the CO 2 concentration <strong>to</strong><br />

300 1000 ppm. Cropping occurs after 10 days,<br />

followed by 2 3 more flushes at weekly<br />

intervals. The whole cycle from compost preparation<br />

<strong>to</strong> cropping is completed in 10 weeks.<br />

Morphogenesis of A. bisporus basidiocarps<br />

The pinheads which form at the surface of the<br />

casing soil attached <strong>to</strong> mycelial cords are<br />

basidiocarp primordia, consisting at first of<br />

loose aggregates of hyphae. Basidiocarp development<br />

has been followed by several workers<br />

(see Bonner et al., 1956; Umar & van Griensven,<br />

1997). The hyphal aggregates become compact<br />

and enlarge <strong>to</strong> form flattened ‘but<strong>to</strong>ns’. When<br />

these are about 2 mm high, there are signs of<br />

reorientation of hyphae in the region of the gills,<br />

and a gap develops beneath the gill region as a<br />

result of programmed cell death (apop<strong>to</strong>sis)<br />

<strong>to</strong> form the hymenial chamber, enclosed below<br />

by the partial veil. At a height of about 5 mm<br />

the stipe becomes recognizable by the parallel<br />

arrangement of hyphae and, by the time the


EUAGARICS CLADE<br />

535<br />

primordium has grown <strong>to</strong> a height of 10 mm, the<br />

orientation of the upper stalk region is complete.<br />

Subsequent increase in height of the stipe is<br />

almost entirely due <strong>to</strong> cell expansion, although<br />

some nuclear and cell division also occur,<br />

especially in the upper region of the stipe<br />

where cell elongation is most marked<br />

(Craig et al., 1977). The gills develop in radially<br />

arranged ridges made up of downward-growing<br />

hyphae terminating in a tightly packed palisade<br />

of cells which are the basidia. Although the<br />

hyphal segments making up the vegetative<br />

mycelium and the stipe are multinucleate,<br />

the number of nuclei in the basidia is reduced<br />

<strong>to</strong> two. As the pileus continues <strong>to</strong> expand, the<br />

partial veil is broken and persists as a ring<br />

attached <strong>to</strong> the stem (Fig. 19.14a). There is<br />

evidence that stipe elongation is promoted by<br />

a fac<strong>to</strong>r of unknown chemical identity, which is<br />

mainly produced by the ripening gills (Frazer,<br />

1996). Konishi (1967) partially purified a<br />

substance which enhanced stipe elongation.<br />

The promoting substance included a mixture of<br />

amino acids, and it is not known if they function<br />

as nutrients or growth fac<strong>to</strong>rs. It has also been<br />

claimed that a compound found chiefly in gill<br />

tissue of mushrooms, 10-oxo-trans-8-decenoic acid<br />

(ODA), stimulates mycelial growth and enhances<br />

elongation of the upper stipe (Mau et al., 1992).<br />

Unfortunately, attempts <strong>to</strong> confirm these findings<br />

have been disappointing (Champavier et al.,<br />

2000).<br />

Biochemical changes occur during basidiocarp<br />

development (Hammond, 1985; de Groot<br />

et al., 1998). Manni<strong>to</strong>l, glycogen and trehalose<br />

accumulate in basidiocarp primordia, lowering<br />

the water potential and thereby possibly causing<br />

uptake of water from the mycelium. This would<br />

lead <strong>to</strong> an increase in turgor pressure, causing<br />

cell enlargement and fruit body expansion<br />

(Hammond & Nichols, 1979). Rapid synthesis<br />

of chitin is correlated with stipe elongation<br />

(Craig et al., 1979).<br />

A hydrophobin-like protein is secreted by the<br />

vegetative mycelium of A. bisporus and an<br />

abundant hydrophobin (ABH1) forms hydrophobic<br />

rodlet layers on the surfaces of the cells in<br />

certain regions of the basidiocarp, especially<br />

those where there are air spaces, such as in the<br />

outer regions of the pileus and stipe, in the veil<br />

and in the core of the stipe, but not in the gills.<br />

The hydrophobic layer may be responsible for the<br />

non-wettability of the surface of the basidiocarp,<br />

preventing the inflow of water from the outside<br />

and possibly protecting against bacterial and<br />

fungal parasites (Lugones et al., 1996).<br />

Life cycle of Agaricus bisporus<br />

The life cycle of A. bisporus is unusual (Miller,<br />

1971). The majority of the spores formed on its<br />

two-spored basidia are at first binucleate but<br />

post-meiotic mi<strong>to</strong>sis increases the number of<br />

nuclei <strong>to</strong> four. The spores are heterokaryotic for<br />

mating type, i.e. they contain non-sister nuclei<br />

with dissimilar A mating type idiomorphs<br />

(Evans, 1959; Elliott, 1985). The products of<br />

meiosis in the basidium are four haploid<br />

nuclei, two with mating type idiomorph A 1 and<br />

two with A 2 . Since two nuclei enter each<br />

basidiospore, the spores may be homokaryotic<br />

(A 1 þ A 1 or A 2 þ A 2 ) or heterokaryotic (A 1 þ A 2 ).<br />

Of the 12 possible pairings of the 4 haploid<br />

meiotic products, 4 are homokaryons and 8 are<br />

heterokaryons, i.e. in a ratio of 1 : 2. Evans (1959)<br />

has claimed that the disproportionately high<br />

ratio of heterokaryotic spores results from the<br />

alignment of the nuclear spindles during meiosis<br />

in the basidia.<br />

On germination the heterokaryotic spores<br />

give rise <strong>to</strong> a mycelium with multinucleate<br />

hyphal segments capable of forming basidiocarps.<br />

Mycelia from homokaryotic spores<br />

can only fruit following anas<strong>to</strong>mosis with mycelia<br />

of opposite mating type. Basidia of normal<br />

cultivated A. bisporus may rarely bear three<br />

or four basidiospores, and most of these are<br />

initially uninucleate. Mycelium from such an<br />

aberrant spore will only fruit when mated with<br />

a mycelium of opposite mating type. Thus<br />

A. bisporus has two alternative kinds of mating<br />

behaviour, secondary homothallism or bipolar<br />

heterothallism. Such ambivalent behaviour is<br />

sometimes termed amphithallic (Lange, 1952)<br />

and is not confined <strong>to</strong> A. bisporus, being<br />

also found in some other basidiomycetes with<br />

two-spored basidia, e.g. certain species of Coprinus<br />

(euagarics clade) and the four-spored Stereum<br />

sanguinolentum (russuloid clade).


536 HOMOBASIDIOMYCETES<br />

Breeding of Agaricus bisporus<br />

The development of new strains of A. bisporus<br />

with superior qualities such as more efficient<br />

substrate conversion rates, rapid fruiting, higher<br />

yield, resistance <strong>to</strong> disease, ease of picking,<br />

extended shelf-life, better appearance, flavour<br />

or consistency is actively pursued by the mushroom<br />

industry (see Raper, 1985; Sonnenberg,<br />

2000). This task is difficult but worthwhile in<br />

view of the high commercial value of the crop,<br />

and various techniques are being used:<br />

1. Selection of strains collected in the field<br />

and arising during cultivation.<br />

2. Hybridization. The small amount of variation<br />

in the gene pool of commercial s<strong>to</strong>cks of<br />

A. bisporus creates little scope for recombination.<br />

However, a search for new genetic resources<br />

within wild Agaricus populations, the Agaricus<br />

Resource Program (ARP), has made available<br />

novel germ plasm which will prove useful in<br />

breeding (Kerrigan, 1996). Of particular interest<br />

is the discovery in the Sonoran desert (California)<br />

of tetrasporic populations, named A. bisporus var.<br />

burnettii, which are completely interfertile with<br />

commercial lines (Kerrigan et al., 1994; Kerrigan,<br />

1995). Agaricus bisporus var. eurotetrasporus,<br />

another four-spored variety, has been discovered<br />

in Europe. Whilst var. burnettii is amphithallic<br />

and predominantly heterothallic, var. eurotetrasporus<br />

is homothallic (Callac et al., 2003).<br />

3. Pro<strong>to</strong>plast fusion. The fusion of pro<strong>to</strong>plasts<br />

from different homokaryons is an alternative <strong>to</strong><br />

the conventional hybridization technique involving<br />

the pairing of compatible homokaryotic<br />

mycelia.<br />

4. Transformation (genetic modification)<br />

using Agrobacterium as a vec<strong>to</strong>r for the transforming<br />

DNA is also possible, but genetically modified<br />

food products are unpopular with consumers<br />

and certain political parties.<br />

Macrolepiota (30 spp.)<br />

The best-known species of Macrolepiota are the<br />

parasol M. procera (Fig. 19.14b) and the shaggy<br />

parasol M. rhacodes, both of which grow in parks,<br />

pastures and woodland. They are saprotrophs.<br />

The pale brown fruit bodies are large and<br />

generally considered good <strong>to</strong> eat, although<br />

those of M. rhacodes may cause gastric upsets.<br />

A feature of both species is that the ring<br />

is detachable and is free <strong>to</strong> move up and<br />

down the stem. Macrolepiota procera and<br />

M. rhacodes have been separated from Lepiota,<br />

and Vellinga et al. (2003) have proposed that they<br />

should be separated from each other as well,<br />

with M. procera showing affinities with<br />

Leucoagaricus and Leucocoprinus, and M. rhacodes<br />

with Agaricus.<br />

There are concerns about the ability of<br />

M. procera <strong>to</strong> concentrate mercury absorbed<br />

from the soil in<strong>to</strong> the basidiocarp tissues, with<br />

values in the caps as high as 13 mgg 1 reported<br />

from various sites in Poland. The mercury<br />

content of the caps and stalks was generally<br />

independent of the soil substrate concentration,<br />

suggesting a remarkable ability of the M. procera<br />

mycelium <strong>to</strong> bioconcentrate mercury (Gucia &<br />

Falandysz, 2003).<br />

Coprinus<br />

Molecular evidence indicates a relationship<br />

between lepio<strong>to</strong>id fungi and two species of<br />

Coprinus, namely C. comatus and C. sterquilinus<br />

(Hopple & Vilgalys, 1999; Redhead et al., 2001).<br />

Although it is now possible, with hindsight, <strong>to</strong><br />

recognize other features shared by these species<br />

and other Agaricaceae, e.g. the shaggy surface of<br />

the cap or the moveable annulus (Fig. 19.14c), we<br />

will discuss them in the context of the many<br />

biological features uniting them with their<br />

former allies (see below).<br />

19.4.2 Coprinaceae (now Psathyrellaceae)<br />

Molecular studies have shown that the genus<br />

Coprinus is not monophyletic (Hopple & Vilgalys,<br />

1999; Redhead, 2001; Redhead et al., 2001),<br />

implying that the coprinoid character of deliquescent<br />

gills has evolved more than once.<br />

This finding has caused a taxonomic nightmare<br />

because one of the two species found <strong>to</strong><br />

be related more closely <strong>to</strong> Macrolepiota than<br />

<strong>to</strong> Coprinus happened <strong>to</strong> be the type-species<br />

C. comatus. Its move from the Coprinaceae <strong>to</strong><br />

the Agaricaceae meant that those numerous<br />

species remaining in the Coprinaceae were<br />

given new names, namely Coprinopsis (e.g. C.<br />

atramentarius, C. cinereus, C. lagopus, C.


EUAGARICS CLADE<br />

537<br />

psychromorbidus), Coprinellus (e.g. C. bisporus, C.<br />

domesticus, C. heptemerus, C. micaceus, C. plagioporus,<br />

C. sassii), and Parasola (e.g. C. plicatilis). Members of<br />

the genus Coprinopsis produce fruit bodies with<br />

hollow stipes and are closely related <strong>to</strong><br />

Psathyrella, thereby presenting the opportunity<br />

<strong>to</strong> re-name Coprinaceae as Psathyrellaceae.<br />

However, because the old name Coprinus in a<br />

broad sense (sensu la<strong>to</strong>) continues <strong>to</strong> be used in<br />

most non-taxonomic studies, we follow that<br />

convention for the time being. It is also possible<br />

that there will eventually be an initiative <strong>to</strong><br />

designate a new type-species of Coprinus, i.e. <strong>to</strong><br />

change the name C. comatus rather than those of<br />

almost all the other species.<br />

Coprinus sensu la<strong>to</strong> (c. 350 spp.)<br />

This is the large group of ‘ink-caps’ with<br />

deliquescent inaequi-hymenial gills undergoing<br />

au<strong>to</strong>digestion at maturity (see Or<strong>to</strong>n & Watling,<br />

1979). Representatives of Coprinus sensu la<strong>to</strong><br />

are cosmopolitan, fruiting on a great variety of<br />

substrata including soil, dung and wood (Or<strong>to</strong>n<br />

& Watling, 1979). Coprinus comatus (the shaggy<br />

ink-cap or lawyer’s wig; Fig. 19.14c) fruits on<br />

soil, especially on disturbed ground. It has a<br />

well-developed, moveable annulus. Another characteristic<br />

feature is the presence of an elastic<br />

strand of aggregated hyphae which extends<br />

through the lumen of the hollow stipe<br />

(Redhead, 2001). Basidiocarps of C. atramentarius<br />

(common ink-cap or pavement cracker) emerge<br />

at the soil surface from buried wood; those of<br />

C. micaceus (glistening ink-cap) are found in<br />

similar situations, or directly on broad-leaved<br />

tree stumps. Coprinus domesticus also fruits on<br />

dead wood of deciduous trees arising from a<br />

rust-coloured sporulating mat of mycelium, the<br />

Ozonium anamorphic state.<br />

There are many coprophilous species, the<br />

most studied being C. cinereus (earlier erroneously<br />

named C. lagopus) on cattle and horse manure<br />

heaps and manured straw (Fig. 19.14d). This<br />

species grows and fruits well in the labora<strong>to</strong>ry<br />

and has been the subject of much research,<br />

ably reviewed by Kües (2000), on the genetics<br />

of mating systems (Pukkila & Cassel<strong>to</strong>n, 1991;<br />

Cassel<strong>to</strong>n & Riquelme, 2004), cy<strong>to</strong>logy (Raju & Lu,<br />

1970; L. Li et al., 1999), morphogenesis (Moore,<br />

1998), gravitropism (Moore et al., 1996), stipe<br />

elongation (Kamada, 1994), nuclear migration,<br />

clamp formation, physiology of fruiting and<br />

other phenomena. Whilst most Coprinus species<br />

are saprotrophic, C. psychromorbidus (¼ ‘diseased<br />

by cold’) belongs <strong>to</strong> a group of low-temperature<br />

basidiomycetes which are plant pathogens. It is<br />

the cause of snow mould disease of grasses and<br />

other plants in Canada and also causes a postharvest<br />

rot of apples in cold s<strong>to</strong>re. It can<br />

continue <strong>to</strong> grow actively under snow cover<br />

because it possesses special thermal hysteresis<br />

proteins with antifreeze properties (Sholberg &<br />

Gaudet, 1992; Hoshino et al., 2003).<br />

Basidiocarps of Coprinus have thin inaequihymenial<br />

gills with prominent pleurocystidia<br />

and cystesia (see p. 525). Basidiocarps range in<br />

size from a few millimetres in coprophilous<br />

forms <strong>to</strong> over 30 cm (C. comatus). Spore development<br />

and discharge occurs first at the base of the<br />

gill, followed by au<strong>to</strong>digestion of the discharged<br />

basidia and the hyphae which supported them.<br />

Chitinase and other hydrolytic enzymes play a<br />

major role in au<strong>to</strong>digestion (Iten & Matile, 1970;<br />

Miyake et al., 1980). The digested tissue drips<br />

away from the base of the gills as a black fluid<br />

which can be used as writing ink. Meanwhile an<br />

upward wave of basidiospore discharge<br />

continues above the digested tissues. In certain<br />

species, the gills do not deliquesce, or do so only<br />

partially. In C. curtus some limited au<strong>to</strong>digestion<br />

occurs at the edge of the gill and the gills open,<br />

as in C. plicatilis, by a V-shaped groove which<br />

widens from above (Buller, 1909, 1931). Buller<br />

(1924) has shown that in many species of Coprinus<br />

the basidia are dimorphic, with long and short<br />

forms present in the same gill. The short forms<br />

are fully functional, i.e. the basidia ripen at<br />

different levels of the hymenium. This arrangement<br />

makes it possible <strong>to</strong> crowd a larger number<br />

of basidia in<strong>to</strong> a given area without interference<br />

in spore release. Trimorphic and tetramorphic<br />

basidia are present in some species.<br />

Sclerotia in Coprinus<br />

Sclerotia are formed by several species. Coprinus<br />

sterquilinus develops its sclerotia at the surface<br />

of cattle dung and, under suitable conditions,<br />

basidiocarps develop from them (Buller, 1924).


538 HOMOBASIDIOMYCETES<br />

In culture, the sclerotia of C. cinereus form on<br />

monokaryons as well as dikaryons growing<br />

aerially or submerged (Waters et al., 1975a,b).<br />

They are quite small (100 1000 mm in diameter),<br />

dark brown <strong>to</strong> black and more or less spherical.<br />

These sclerotia do not develop basidiocarps.<br />

Three tissue layers are distinguishable in aerial<br />

sclerotia: an outer diffuse layer of apparently<br />

dead cells, a multi-layered rind of heavily<br />

pigmented, closely packed, thick-walled cells,<br />

and a medulla with predominantly thick-walled<br />

cells. These thick-walled cells at first accumulate<br />

rosettes of a glycogen-like polysaccharide in the<br />

cy<strong>to</strong>plasm, but this disappears as the cells<br />

mature, while a secondary cell wall layer<br />

increases in thickness. It is believed that the<br />

glycogen serves a temporary s<strong>to</strong>rage function but<br />

that the secondary wall represents the long-term<br />

s<strong>to</strong>rage compartment.<br />

Stipe elongation in Coprinus<br />

Shortly before spore discharge begins in Coprinus,<br />

the stipe undergoes rapid elongation, most of it<br />

taking place in its upper half. Although in<br />

C. cinereus the cells of the vegetative mycelium<br />

are dikaryotic, the cortical cells of the stipe<br />

undergo repeated nuclear division prior <strong>to</strong> rapid<br />

elongation so that they may contain between<br />

32 and 156 nuclei. Stipe elongation is largely<br />

brought about by turgor-driven cell expansion,<br />

not cell division. The cells increase in length, but<br />

not in width, by about eightfold in the final 15 h<br />

of fruit body expansion. During that period, the<br />

osmotic pressure is actively maintained as water<br />

enters the cells. The solutes contributing <strong>to</strong> the<br />

maintenance of osmotic pressure have not been<br />

identified. The increase in stipe length is<br />

mirrored by an increase in the chitin content.<br />

Helically coiled chitin microfibrils are present in<br />

the walls of cortical cells, and it is believed that<br />

newly synthesized microfibrils are inserted<br />

between pre-existing ones, a process termed<br />

diffuse extension growth <strong>to</strong> distinguish it from<br />

the usual apical extension growth of hyphal tips<br />

(Kamada, 1994).<br />

The force of the elongating stipe can be<br />

considerable. Buller (1931) placed weights on an<br />

expanding fruit body of C. sterquilinus which had<br />

a pileus height of 1.4 cm and diameter of 0.8 cm.<br />

This fruit body could lift a weight of 204 g.<br />

The cross-sectional area of the ring of tissue<br />

surrounding the hollow stipe was 29 mm 2<br />

and calculations of the upward pressure of the<br />

stipe gave a value of two-thirds of an atmosphere<br />

(0.7 10 4 Nm 2 ). This may explain why the<br />

expanding fruit bodies of C. atramentarius can<br />

crack asphalt and why other fungi can lift paving<br />

slabs.<br />

Mating behaviour in Coprinus<br />

Some Coprinus spp. are primarily homothallic,<br />

e.g. C. heptemerus. Species with two-spored basidia<br />

are often secondarily homothallic, e.g. C. bisporus<br />

(¼ C. ephemerus var. bisporus). Yet other species<br />

show bipolar (i.e. unifac<strong>to</strong>rial) heterothallism<br />

(e.g. C. comatus and C. ephemerus), or tetrapolar<br />

(i.e. bifac<strong>to</strong>rial) heterothallism (e.g. C. cinereus;<br />

see Section 18.9). M. Lange (1952) has proposed<br />

the term amphithallism (see p. 532) <strong>to</strong> describe<br />

the behaviour of species of Coprinus in which<br />

both homothallic and heterothallic mycelia can<br />

be raised from the same fruit body, giving as<br />

examples C. sassii, a bisporic species which is<br />

amphithallic-bipolar, and the amphithallictetrapolar<br />

C. plagioporus.<br />

In C. cinereus, as in many other species,<br />

monokaryotic (rarely dikaryotic) mycelia form<br />

globose mucilaginous heads containing numerous<br />

oidia dispersed by insects. Plasmogamy<br />

preceding dikaryotization may be achieved<br />

by fusion between compatible monokaryotic<br />

hyphae or by fusion between a monokaryotic<br />

hypha and a compatible oidium. Monokaryons<br />

may also be dikaryotized following hyphal fusion<br />

with dikaryons (the Buller phenomenon; see<br />

pp. 508 and 566).<br />

Edibility of Coprinus basidiocarps<br />

Some species of Coprinus have basidiocarps which<br />

are edible when young. Coprinus comatus is an<br />

example. The fruit bodies of C. atramentarius are<br />

edible and harmless except if consumed with<br />

alcohol, which results in unpleasant symp<strong>to</strong>ms<br />

of nausea and palpitations. The substance associated<br />

with this effect has been identified as<br />

coprine (Fig. 19.15a), and this is similar in its<br />

effects <strong>to</strong> the drug disulfuram (antabuse) which


EUAGARICS CLADE<br />

539<br />

Fig19.15 Secondary metabolites in the euagarics clade.Coprine (a) produced by Coprinus atramentarius is <strong>to</strong>xic only when consumed with alcohol. a-Amanitin (b) and phalloidin (c) are the<br />

<strong>to</strong>xic principles of death caps, especially Amanita phalloides and A. virosa.The alkaloids ibotenic acid (d) and its derivative muscimol (e) are found in the fly agaric, Amanitamuscaria.Theyare<br />

hallucinogenic, probably mimicking neurotransmitters. Similarly, the indole alkaloids psilocybin (f) and its derivative psilocin (g), produced by Psilocybe spp., are hallucinogenic.Their<br />

structure is similar <strong>to</strong> the neurotransmitter sero<strong>to</strong>nin.The bipyridyl <strong>to</strong>xin orellanine (h) is the cause of poisoning by Cortinarius orellanus.


540 HOMOBASIDIOMYCETES<br />

is used in attempts <strong>to</strong> wean alcoholics from their<br />

addiction, although it is different chemically.<br />

Interference competition involving Coprinus<br />

Many species of Coprinus are coprophilous, typically<br />

with their basidiocarps appearing relatively<br />

late in the succession of fungi on dung (Dix &<br />

Webster, 1995). Certain species are known <strong>to</strong><br />

suppress the fruiting of other fungi. A good<br />

example of this phenomenon is C. heptemerus,<br />

which inhibits the fruiting of many species in<br />

nature on rabbit dung and in culture. Its effect<br />

on the sensitive ascomycete Ascobolus crenulatus<br />

was tracked down <strong>to</strong> the moment when the<br />

hyphae of the two fungi make contact. Within<br />

minutes, the hyphal segment of Ascobolus<br />

<strong>to</strong>uched by a hyphal tip of Coprinus is killed.<br />

There is a rapid loss of turgor, shown by the<br />

bulging of the septa of adjacent cells in<strong>to</strong> the<br />

affected cell, which also loses the ability <strong>to</strong><br />

undergo plasmolysis, whilst adjacent cells readily<br />

plasmolyse when bathed in hyper<strong>to</strong>nic fluid.<br />

This form of competition has been termed<br />

hyphal interference or interference competition.<br />

It occurs in a range of genera of coprophilous<br />

basidiomycetes, but is not confined <strong>to</strong> this ecological<br />

group, having been demonstrated <strong>to</strong> be<br />

effective also in lignicolous fungi (Ikediugwu &<br />

Webster, 1970a,b; Ikediugwu et al., 1970).<br />

Cell damage is very similar <strong>to</strong> the effects seen<br />

when self- and non-self hyphae of the same<br />

species confront each other. An oxidative burst is<br />

stimulated and hydrogen peroxide accumulates<br />

in cells of the sensitive partner. This is interpreted<br />

as a defence reaction (Silar, 2005).<br />

19.4.3 Amanitaceae<br />

Amanita (c. 500 spp.)<br />

This is a large and important genus whose<br />

species form sheathing mycorrhiza with trees.<br />

Amanita muscaria (‘fly agaric’; Plate 9a) is often<br />

associated with birch (Betula) but also grows in<br />

mycorrhizal association with Abies, Pinus, Picea,<br />

Quercus and other hosts. As is well known,<br />

the basidiocarps of some species are poisonous,<br />

especially those of A. phalloides (death cap),<br />

A. virosa (destroying angel), A. pantherina (panther<br />

cap) and A. verna, whilst those of A. muscaria are<br />

more hallucinogenic than poisonous. There are<br />

also species whose basidiocarps are excellent <strong>to</strong><br />

eat, most notably A. caesarea (Caesar’s mushroom;<br />

Plate 9b), which has been hunted enthusiastically<br />

in Southern European countries since Roman<br />

times. Emperor Claudius was an early connoisseur<br />

of A. caesarea and may have paid for his<br />

mycophagy with his life, probably falling victim<br />

<strong>to</strong> a poisoned mushroom dish manipulated by<br />

his wife Agrippina in AD 54 (Ramsbot<strong>to</strong>m, 1953).<br />

Other edible species are A. rubescens (blusher),<br />

A. vaginata (grisette) and A. fulva (tawny grisette).<br />

In view of the possible confusion between edible<br />

and poisonous species, it is obviously best <strong>to</strong><br />

avoid eating basidiocarps of any whose identity<br />

is uncertain.<br />

The characteristic features of Amanita include<br />

a white spore print and the presence of a volva,<br />

i.e. the <strong>to</strong>rn remnants of a universal veil. The<br />

volva persists as a cup at the base of the stipe and<br />

broken volva fragments may also adhere <strong>to</strong> the<br />

cap, as seen as the white scales on the red caps of<br />

A. muscaria (Plate 9a). Most species also have<br />

a ring (annulus) on the stem, the remnant of the<br />

partial veil which protected the gills during fruit<br />

body development, but there is no ring in<br />

some species formerly classified in Amani<strong>to</strong>psis<br />

(e.g. A. vaginata and A. fulva).<br />

Amanita poisoning<br />

Symp<strong>to</strong>ms after ingestion of fruit bodies of<br />

A. phalloides follow a characteristic time course<br />

over a period of 7 days (Faulstich & Zilker, 1994).<br />

After a mushroom meal and symp<strong>to</strong>m-free<br />

interval (day 1), there is a period of emesis<br />

(vomiting), abdominal cramps and diarrhoea<br />

(day 2), followed by a period of remission (day<br />

3) which is treacherous because it lures many<br />

patients in<strong>to</strong> believing that they have overcome<br />

the poisoning. Meanwhile, severe liver damage is<br />

ongoing and symp<strong>to</strong>ms resume with a vengeance<br />

with gastrointestinal bleeding (day 4), hepatic<br />

encephalopathy (brain damage, day 5), kidney<br />

failure (day 6) and death (day 7).<br />

Amanita poisoning is caused by two <strong>to</strong>xins,<br />

namely the ama<strong>to</strong>xin a-amanitin and the<br />

phallo<strong>to</strong>xin phalloidin (Bresinsky & Besl, 1990;<br />

Wieland & Faulstich, 1991; Chil<strong>to</strong>n, 1994,<br />

Wieland, 1996). Both are bicyclic oligopeptides


EUAGARICS CLADE<br />

541<br />

(see Figs. 19.15b,c). Of these, the more damaging<br />

is a-amanitin which binds <strong>to</strong> hepa<strong>to</strong>cytes (liver<br />

cells). It inhibits a nuclear polymerase responsible<br />

for transcribing DNA in<strong>to</strong> mRNA, resulting<br />

in reduced protein synthesis at the ribosomes<br />

and ultimately the death of cells. Phalloidin acts<br />

by binding <strong>to</strong> G-actin in liver cells and brings<br />

about an efflux of potassium ions and lysosomal<br />

enzymes, which leads <strong>to</strong> cell destruction<br />

(Chil<strong>to</strong>n, 1994).<br />

The treatment of patients suffering from<br />

A. phalloides poisoning includes forced vomiting<br />

and other means of evacuating the s<strong>to</strong>mach,<br />

orally applied activated charcoal <strong>to</strong> adsorb the<br />

<strong>to</strong>xins, external dialysis, infusion with silybinin,<br />

an extract from the milkthistle Silybum marianum<br />

which competes with a-amanitin for adsorption<br />

on <strong>to</strong> hepa<strong>to</strong>cytes, and liver transplantation<br />

(Faulstich & Zilker, 1994).<br />

The effects of ingesting basidiocarps of<br />

A. muscaria are less severe, being mainly hallucinogenic.<br />

Characteristic symp<strong>to</strong>ms are drowsiness,<br />

followed by deep sleep in which vivid<br />

dreams occur. Recovery is generally complete,<br />

with no permanent or prolonged ill after-effects.<br />

The two <strong>to</strong>xins chiefly involved are the alkaloids<br />

ibotenic acid and muscimol (Figs. 19.15d,e;<br />

Michelot & Melendez-Howell, 2003), and the<br />

former is readily converted <strong>to</strong> the latter in the<br />

gut. The molecular structures of ibotenic acid<br />

and muscimol closely resemble those of two<br />

neurotransmitters, glutamic acid and g-aminobutyric<br />

acid (GABA), respectively. Certain ethnic<br />

groups (e.g. in Siberia) have taken advantage of<br />

the hallucinogenic properties of A. muscaria <strong>to</strong><br />

experience euphoria. Its use has extended <strong>to</strong><br />

semi-religious practices in which shamans have<br />

induced themselves in<strong>to</strong> trances in which they<br />

claim <strong>to</strong> have powers of revelation. Within 1 h of<br />

ingestion, ibotenic acid and muscimol are<br />

detectable in the urine of humans and also<br />

reindeer, and this may account for the tradition<br />

of drinking such urine in order <strong>to</strong> obtain a<br />

‘second-hand kick’. Such practices have now been<br />

discontinued in favour of alcoholic beverages,<br />

although A. muscaria is still occasionally taken as<br />

a recreational drug.<br />

The name ‘fly agaric’ refers <strong>to</strong> the supposed<br />

insecticidal properties of A. muscaria caps. When<br />

soaked in milk the caps attract flies, possibly<br />

brought there in response <strong>to</strong> an attractant,<br />

diolein. The flies ingest the Amanita flesh, but it<br />

is likely that they are in<strong>to</strong>xicated rather than<br />

killed.<br />

Amanita muscaria, easily recognized by its red<br />

cap adorned by white volva scales, is the bestknown<br />

of all macro-fungi, featuring in countless<br />

illustrations, folk tales, religious and semireligious<br />

ceremonies. It has been considered as<br />

the ‘tree of life’ or ‘tree of knowledge’ (Wasson,<br />

1968). Michelot and Melendez-Howell (2003) have<br />

given an account of its chemistry, biology,<br />

<strong>to</strong>xicology and ethnomycology. It is widely<br />

distributed in Europe, North America and Asia,<br />

and has probably been introduced in<strong>to</strong> other<br />

parts of the world as a mycorrhizal partner on<br />

the roots of imported trees. Molecular phylogeny<br />

studies suggest that it should be separated in<strong>to</strong><br />

at least three groups, corresponding <strong>to</strong> Eurasian,<br />

Eurasian subalpine and North American regions<br />

(Oda et al., 2004).<br />

19.4.4 Pluteaceae<br />

Pluteus (c. 300 spp.)<br />

Species of Pluteus are saprotrophs, growing<br />

mainly on rotting wood. Pluteus cervinus is<br />

common, fruiting on deciduous tree stumps,<br />

logs, fallen branches and on sawdust heaps.<br />

Characteristic features are the free gills and<br />

pink spore print. On the surface of the hymenium,<br />

tapering, thick-walled cystidia crowned by<br />

pointed prongs protrude (Fig. 19.9d). Their function<br />

is unknown.<br />

The Pluteaceae are related <strong>to</strong> the<br />

Amanitaceae, and these two families appear as<br />

sister groups in phylogenetic analyses (Moncalvo<br />

et al., 2002).<br />

19.4.5 Pleurotaceae<br />

This family includes only two genera, Pleurotus<br />

and Hohenbuehelia (Thorn et al., 2000; Moncalvo<br />

et al., 2002).<br />

Pleurotus (20 spp.)<br />

Pleurotus spp. (oyster mushrooms) have laterally<br />

attached, lamellate basidiocarps. Most cause<br />

white rot decay of wood. The basidiocarps are


542 HOMOBASIDIOMYCETES<br />

edible and several species, e.g. P. ostreatus and<br />

P. sajor-caju, are widely cultivated. Ligno-cellulosic<br />

waste products from a range of agricultural<br />

crops (e.g. wheat, rice, maize straw, banana<br />

leaves and dried water hyacinth) and industrial<br />

extraction processes can be used as substratum<br />

and cultivation can be practised in fac<strong>to</strong>ries as<br />

well as on a smaller scale, e.g. in large cylindrical<br />

plastic bags in the context of a ‘cottage industry’<br />

which can provide a valuable food supplement<br />

in developing countries (Poppe, 2000; Sánchez,<br />

2004). Spent substratum can be used for<br />

further fungal fermentations or as animal<br />

feed. A problem with the cultivation of Pleurotus<br />

spp. is the massive amount of basidiospores<br />

released from an early developmental stage<br />

onwards, causing ‘mushroom worker’s lung’.<br />

In order <strong>to</strong> overcome this problem, sporeless<br />

mutants have been bred.<br />

Pleurotus ostreatus, one of the best-known<br />

species, forms clusters of fan-shaped, bluishgrey<br />

basidiocarps at the base of deciduous tree<br />

stumps (Fig. 19.14e), especially beech (Fagus<br />

sylvatica). The gills are decurrent (running down<br />

the base of the stipe). Development in P. ostreatus<br />

is gymnocarpic and construction monomitic, but<br />

in some other species, e.g. P. tuberregium, skeletal<br />

hyphae are present. Some species develop<br />

thallic arthric anamorphs, e.g. P. cystidiosus in<br />

which the conidiophores are synnematal. In<br />

addition <strong>to</strong> being able <strong>to</strong> decompose wood,<br />

P. ostreatus and some related species are nema<strong>to</strong>phagous<br />

(Barron & Thorn, 1987; Hibbett & Thorn,<br />

1994; Thorn et al., 2000; see p. 679), and the<br />

nema<strong>to</strong>de prey represents an important supplement<br />

of nitrogen which is often the growthlimiting<br />

nutrient in woody substrata. The<br />

ability <strong>to</strong> parasitize nema<strong>to</strong>des has been used<br />

as a criterion <strong>to</strong> support the classification of<br />

P. tuberregium, a tropical species whose basidiocarps<br />

arise from large subterranean sclerotia<br />

(Hibbett & Thorn, 1994).<br />

Hohenbuehelia (50 spp.)<br />

This genus includes terrestrial lignicolous<br />

species. The basidiocarps frequently contain<br />

gelatinized tissue. All species are nema<strong>to</strong>phagous.<br />

Nema<strong>to</strong>des are trapped on hourglassshaped<br />

(i.e. constricted) knobs covered with a<br />

<strong>to</strong>xic secretion, formed either directly on the<br />

clamped mycelium or at the tips of bent, tapering<br />

conidia assigned <strong>to</strong> the anamorph genus<br />

Nema<strong>to</strong>c<strong>to</strong>nus (Barron, 1977; Barron & Dierkes,<br />

1977; see Fig. 25.7).<br />

19.4.6 Schizophyllaceae<br />

Although the unique pattern of fruit body and<br />

gill morphogenesis seemed <strong>to</strong> set Schizophyllum<br />

clearly apart from the Agaricales (Donk, 1964),<br />

the euagarics clade is where we must now<br />

place it on the basis of DNA analyses (Moncalvo<br />

et al., 2002). More confusingly still, the two fungi<br />

apparently closest <strong>to</strong> Schizophyllum are the<br />

polypore Fistulina and the gill-bearing genus<br />

Volvariella.<br />

Schizophyllum<br />

Schizophyllum commune has a worldwide distribution<br />

but is more common in warmer regions.<br />

It grows saprotrophically or parasitically on a<br />

wide range of woody substrata forming beigecoloured,<br />

fan-shaped, laterally attached fruit<br />

bodies with a furry upper surface. Since 1990 it<br />

has become common in Western Europe on<br />

plastic-wrapped hay silage bales from which<br />

clusters of basidiocarps burst out (Fig. 19.16a;<br />

Brady et al., 2005). Its spores are wind-borne.<br />

James and Vilgalys (2001), working in the<br />

Caribbean, trapped spores sedimenting on<strong>to</strong><br />

the surface of Petri dishes containing a pregrown<br />

homokaryon culture. A successful trapping<br />

event was detected by the appearance of<br />

dikaryotic growth (with clamp connections).<br />

Deposition rates of 18 spores m 2 h 1 indicate<br />

that there is ample inoculum <strong>to</strong> ensure the<br />

colonization of most available substrates as soon<br />

as they are available for decay.<br />

The name Schizophyllum refers <strong>to</strong> the longitudinally<br />

‘split gills’ which are a xeromorphic<br />

adaptation (Figs. 19.16b and 19.17). In dry<br />

weather the ‘gills’ curve inwards so that the<br />

hymenial surface is protected by a series of<br />

adjoining folds. The curvature is due <strong>to</strong> the<br />

shrinkage of thinner-walled hyphal layers on<br />

drying. Since the remaining tissue of the gill is<br />

composed of thick-walled clamped hyphae which<br />

do not shrink so readily, inward curvature<br />

follows. Dried basidiocarps can be s<strong>to</strong>red for


EUAGARICS CLADE<br />

543<br />

Fig19.16 Basidiocarps in the euagarics clade (2). (a,b) Schizophyllum commune. (a) Cluster of fruit bodies growing through the plastic<br />

covering of a silage bale. (b) Basidiocarp as seen from below. (c) Volvariella speciosa. Note the presence of a volva and the absence<br />

of a ring. (d) Volvariella surrecta, a mycoparasite fruiting on basidiocarps of Cli<strong>to</strong>cybe nebularis.(e)Panaeolus sphinctrinus.<br />

years and revive when placed in a moist<br />

environment. Water uptake occurs especially<br />

through the hairy upper surface, and the gills<br />

straighten out as the hymenium expands<br />

within 2 3 h. Spore discharge commences after<br />

3 4 h, and basidia are readily seen in hand<br />

sections, rendering S. commune a suitable object<br />

for microscopic examination if classes have <strong>to</strong> be<br />

held out of season.<br />

Schizophyllum commune is a ‘model’ fungus<br />

with a very extensive literature. It has been<br />

studied in detail by workers interested in its<br />

mating system, which is tetrapolar with complex<br />

A and B loci, each composed of two sub-loci,<br />

a and b. Several alleles have been identified at<br />

each sub-locus, creating over 20 000 mating<br />

type specificities (see p. 506). Other aspects of<br />

its biology which have been studied in depth


544 HOMOBASIDIOMYCETES<br />

Fig19.17 Schizophyllum commune. (a) V.S. portion of<br />

basidiocarp showing the divided inrolled ‘gills’.<br />

(b) High-power drawing of part of the‘gill’ in the<br />

region of the split (arrowed). Note that the hyphae<br />

in this region are thin-walled in contrast <strong>to</strong> the<br />

thicker-walled hyphae making up the rest of the<br />

flesh.<br />

include genetics (Raper & Hoffman, 1974),<br />

nuclear migration (Snider, 1965) and morphogenesis<br />

(Wessels, 1993b, 1994). Dikaryotic mycelia<br />

fruit readily in culture. Monokaryotic fruiting<br />

may also occur (for references see Moore, 1998).<br />

The early stages of basidiocarp development have<br />

been studied by Raudaskoski and Viitanen (1982)<br />

and Raudaskoski and Vauras (1982). A short<br />

exposure <strong>to</strong> light and good aeration are necessary<br />

<strong>to</strong> induce sporulation; elevated CO 2 concentrations<br />

are inhibi<strong>to</strong>ry. Under dark conditions,<br />

mycelial growth is depressed beneath the surface<br />

of the medium under a layer of slimy material.<br />

Within 3 h after transfer of 3 day old dark-grown<br />

cultures <strong>to</strong> continuous light, there is abundant<br />

growth of aerial hyphae at the margin of the<br />

culture. After 9 h the emergent aerial hyphae<br />

become aggregated in<strong>to</strong> a horseshoe-shaped<br />

structure which is the ventral surface of the<br />

fruit body. After 15 h basidial initials form the<br />

beginnings of the hymenium which develops<br />

over the entire lower surface of the basidiocarp.<br />

The outside of the fruit body consists of vertical<br />

and horizontal hyphal strands, and its outermost<br />

layer is formed by parallel backward growth of<br />

numerous hyphal tips <strong>to</strong>wards the developing<br />

hymenium. By more rapid growth on one side,<br />

the fruit body may become fan-shaped. The split<br />

gills arise by marginal proliferation, and their<br />

number is increased by downgrowths from the


EUAGARICS CLADE<br />

545<br />

flesh of the fruit body. The cy<strong>to</strong>logy of basidial<br />

development shows no unusual features.<br />

Surface properties, especially the wettability<br />

of hyphal surfaces, play an important role in<br />

mycelial growth and fruit body development in<br />

Schizophyllum. Aerial hyphae, including those<br />

making up the basidiocarp, are strongly hydrophobic<br />

(i.e. non-wettable) because they are<br />

covered by parallel rodlets of special proteins,<br />

hydrophobins, which were first discovered<br />

during research with S. commune. Hydrophobins<br />

are amphipathic, i.e. they arrange themselves<br />

in<strong>to</strong> sheets with one wettable (hydrophilic)<br />

face and a non-wettable (hydrophobic) face.<br />

At an air water interface or over the surface<br />

of a hypha, the hydrophilic face attaches <strong>to</strong> the<br />

meniscus or <strong>to</strong> the hyphal surface whilst the<br />

hydrophobic layer faces outwards, giving it its<br />

non-wettable properties (Wessels, 1997, 2000).<br />

Schizophyllum commune has at least four hydrophobin<br />

genes. The gene SC3 is active in both<br />

monokaryons and dikaryons, whereas SC1, SC4<br />

and SC6 are active in dikaryons only. When<br />

grown experimentally in liquid culture, the<br />

hyphae of S. commune secrete the monomers of<br />

SC3p which accumulate at the air liquid interface<br />

and also lower the surface tension of<br />

the culture liquid. This enables individual<br />

hyphae <strong>to</strong> penetrate the interface, and <strong>to</strong> grow<br />

in<strong>to</strong> the air, simultaneously becoming coated<br />

with hydrophobin. The function of the product<br />

of the SC4 gene, the hydrophobin SC4p, has been<br />

clarified. It covers the surface of the hyphae<br />

lining the numerous air channels which traverse<br />

the basidiocarp, preventing them from being<br />

clogged by water and thus permitting gas<br />

exchange <strong>to</strong> continue during respiration.<br />

Volvariella<br />

There are about 50 species of Volvariella, and their<br />

taxonomic position is still unsettled but may<br />

be close <strong>to</strong> Schizophyllum (Moncalvo et al., 2002).<br />

The genus includes several species cultivated<br />

for their edible basidiocarps. The best known is<br />

V. volvacea, the paddy straw mushroom, widely<br />

grown in Asia on rice straw, cot<strong>to</strong>n waste and<br />

other cellulose-rich agricultural waste products.<br />

In terms of world production it is one of<br />

the most important edible mushrooms. It is a<br />

warm-temperature fungus which can grow vegetatively<br />

at 32 34°C, with an optimum temperature<br />

for fruiting of 28 30°C. Under favourable<br />

conditions, the period between inoculation and<br />

harvest of fruit bodies is 8 10 d, the shortest in<br />

any cultivated fungus (Chang & Miles, 2004).<br />

The basidiocarps of Volvariella spp. are enclosed<br />

by a universal veil which persists as a prominent<br />

cup-like volva (Fig. 19.16c). There is no ring.<br />

The gills are free and the spore print is pink.<br />

Volvariella volvacea has an unusual life cycle<br />

(Chang & Yau, 1971; Chiu, 1993). It is homothallic<br />

and its mycelium is haploid, being made up of<br />

multinucleate segments. There are no clamp<br />

connections and it is reported that nuclei can<br />

pass through the transverse septa which separate<br />

adjacent hyphal segments. Brown, thick-walled,<br />

multinucleate chlamydospores are borne on<br />

specialized branches of the aerial mycelium.<br />

They serve as asexual propagules under adverse<br />

conditions, germinating by hyphal growth.<br />

Young basidia contain two haploid nuclei<br />

which fuse <strong>to</strong> form a diploid nucleus. Meiosis<br />

gives rise <strong>to</strong> four haploid nuclei, one entering<br />

each of the four basidiospores. Basidiospores are<br />

therefore normally uninucleate (occasionally<br />

binucleate).<br />

Volvariella bombycina, the silver silk straw<br />

mushroom, has a similar life cycle. It fruits<br />

readily in labora<strong>to</strong>ry culture and its basidiocarp<br />

development has been studied by Chiu and<br />

Moore (1990b). Development is normally hemiangiocarpic,<br />

but angiocarpic (i.e. with the<br />

hymenium enclosed until a late stage) and<br />

gymnocarpic (hymenium exposed) development<br />

may also occur. Hemi-angiocarpic development<br />

has been arbitrarily divided in<strong>to</strong> five<br />

phases: the bulb, but<strong>to</strong>n, egg, elongation and<br />

mature stages. During the but<strong>to</strong>n stage, a<br />

schizogenous cavity is formed, enclosed by the<br />

universal veil and basal bulb. The hymenophore<br />

develops within this cavity. Elongated cylindrical<br />

hyphae of uniform size develop on ridges,<br />

initiating the formation of the gills. These<br />

ridges do not extend <strong>to</strong> the future stipe, so the<br />

gills remain free. During the late egg stage, i.e.<br />

before rupture of the universal veil, local<br />

proliferation of the gills occurs, and at the<br />

site of proliferation the hymenophore is folded


546 HOMOBASIDIOMYCETES<br />

in a sinuous or labyrinth-like pattern in contrast<br />

with the regular folding seen in the mature<br />

stage. During the elongation stage, expansion of<br />

the stipe and cap causes the universal veil <strong>to</strong><br />

rupture. In addition <strong>to</strong> the primary gills which<br />

run along the entire radius of the cap, secondary<br />

or tertiary gills, extending for shorter distances,<br />

develop either by successive bifurcation of an<br />

older gill near its inner, free edge or by the<br />

formation of a new hymenial layer at or near the<br />

root of an old gill. The hymenium includes<br />

basidia and larger, skittle-shaped facial cystidia<br />

which may extend across <strong>to</strong> the opposite gill<br />

face. Droplets observed on the cystidia suggest<br />

that the cystidia may have a secre<strong>to</strong>ry role.<br />

Volvariella surrecta is mycoparasitic. Its<br />

basidiocarps grow on the caps of other agarics<br />

such as Cli<strong>to</strong>cybe nebularis (Fig. 19.16d).<br />

19.4.7 Bolbitiaceae<br />

Panaeolus (25 spp.)<br />

Most members are coprophilous but some grow<br />

in pastures. Common examples are P. semi-ovatus<br />

(sometimes classified in a separate genus<br />

Anellaria) and P. sphinctrinus (Fig. 19.16e). Both<br />

species fruit on cattle dung. The gills are<br />

wedge-shaped in section and aequi-hymeniiferous.<br />

The spores are black. Panaeolus spp. are<br />

commonly called mottle-gills, and this refers <strong>to</strong><br />

the fact that the basidia ripen in patches, not<br />

uniformly, so that areas in which the spores<br />

are ripe appear darker than those in which<br />

the spores are still immature (Buller, 1922).<br />

Basidiocarps of some Panaeolus spp. contain<br />

psilocybin and are hallucinogenic (see p. 553).<br />

19.4.8 Hygrophoraceae<br />

Members of the Hygrophoraceae are sometimes<br />

included in Tricholomataceae but have been<br />

resolved in recent phylogenetic studies as a<br />

separate family (Moncalvo et al., 2002).<br />

Representative genera are Hygrophorus (100 spp.)<br />

and Hygrocybe (150 spp.), also known as waxcaps<br />

because of the texture of their basidiocarps.<br />

Some species grow in woodland and may form<br />

ec<strong>to</strong>mycorrhiza with trees, e.g. Hygrophorus<br />

eburneus from beech woods (Fagus sylvatica).<br />

However, the characteristic habitats of many<br />

waxcaps are impoverished, non-fertilized<br />

pastures. Their basidiocarps are often slimy or<br />

gelatinous and brightly coloured, e.g. the bloodred<br />

Hygrocybe coccinea (scarlet waxcap; Plate 9c) or<br />

the greenish-yellow Hygrocybe psittacina (parrot<br />

waxcap). The basidiocarps of Hygrocybe conica<br />

(blackening waxcap) turn black when bruised.<br />

19.4.9 Marasmiaceae<br />

Although the Marasmiaceae, like most families<br />

in the euagarics clade, may well become substantially<br />

rearranged in the future, the work by<br />

Moncalvo et al. (2000) lends some support for<br />

the grouping <strong>to</strong>gether of key genera such as<br />

Lentinula, Omphalotus, Marasmius, Crinipellis,<br />

Flammulina and Strobilurus. Most fungi formerly<br />

known as Collybia are now also accommodated<br />

in the Marasmiaceae, assigned <strong>to</strong> the genera<br />

Gymnopus and Rhodocollybia. Several members<br />

of the Marasmiaceae are known <strong>to</strong> produce<br />

antifungal substances in pure culture, and one<br />

group, the strobilurins, has been developed in<strong>to</strong><br />

fungicides currently enjoying worldwide application<br />

(Sauter et al., 1999; see p. 410). Some<br />

strobilurins and the related oudemansins have<br />

been shown <strong>to</strong> be secreted by their producers<br />

in<strong>to</strong> colonized wood at concentrations which are<br />

<strong>to</strong>xic <strong>to</strong> potential competi<strong>to</strong>rs, thereby indicating<br />

their ecophysiological function as a means of<br />

resource capture and defence (Engler et al., 1998).<br />

Whereas most members of the Marasmiaceae are<br />

saprotrophs on wood and humus, some are<br />

necrotrophic or, rarely, biotrophic pathogens of<br />

trees.<br />

Marasmius (500 spp.)<br />

Species of Marasmius often have rather <strong>to</strong>ugh,<br />

leathery basidiocarps which shrivel on drying<br />

but rapidly revive on wetting. The best known<br />

species is M. oreades, the fairy ring fungus whose<br />

basidiocarps are edible and may be dried<br />

(Fig. 19.18a). The mycelium of the fungus<br />

grows outwards from a central point and, as<br />

the mycelial front progresses radially, the<br />

older, trailing mycelium dies. This results in a<br />

ring of active mycelium, visible as a circle of<br />

green grass in lawns and pastures (Fig. 19.18b).<br />

Measurements of the radial extension of growing


EUAGARICS CLADE<br />

547<br />

Fig19.18 Basidiocarps in the euagarics clade (3): Marasmiaceae. (a) Marasmius oreades.(b)TwofairyringsofM. oreades approaching<br />

each other on a lawn.The centres of the rings are <strong>to</strong> the <strong>to</strong>p and bot<strong>to</strong>m of the picture. (c) Oudemansiella radicata which grows<br />

attached <strong>to</strong> tree stumps andburied wood andhas a long, tapering, undergroundbase <strong>to</strong> the stem (pseudorhiza). (d) Armillaria mellea.<br />

(e) Flower cushion of cocoa tree infected by the monokaryotic mycelium of Crinipellis perniciosa.Bothhealthyflowersandswollen<br />

infected shoots are visible. (f) Basidiocarps of C. perniciosa on a dead cocoa twig. (e) and (f) from Griffith et al.(2003),New Zealand<br />

Journal of Botany, by copyright permission of the Royal Society of New Zealand; original images kindly provided by R.N. Birch, J.N.<br />

Hedger and G.W.Griffith.<br />

rings show that they can extend outwards at a<br />

rate of about 1 3.5 cm per annum, and measurements<br />

of their diameter indicate that some,<br />

in permanent pastures, may be centuries old.<br />

The mycelium is intermingled with the grass<br />

roots usually near the soil surface at a depth of<br />

8 10 cm, but it may extend <strong>to</strong> a depth of 30 cm,<br />

depending on soil type. Closer inspection shows<br />

that there are three concentric rings at the soil<br />

surface: (1) an outer zone in which the grass is<br />

greener and taller than outside, and in which the<br />

fungus fruits; (2) a middle zone where the<br />

grass is dead and the ground bare, especially in<br />

dry seasons; and (3) an inner zone of stimulated


548 HOMOBASIDIOMYCETES<br />

growth often occupied by other plants which<br />

have colonized the previously bare ground.<br />

The stimulated growth of grass in the outer<br />

ring zone is associated with more rapid decomposition<br />

of soil organic matter and the concomitant<br />

release of nutrients. The amount of soil<br />

organic carbon may be lowered by about 50% as<br />

compared with a non-invaded turf. The death of<br />

grass in the bare central ring zone is probably<br />

due <strong>to</strong> a combination of fac<strong>to</strong>rs, including<br />

parasitic attack by the fungus, drought resulting<br />

from impeded water percolation, and <strong>to</strong>xins of<br />

fungal origin (e.g. cyanide, HCN) which damage<br />

the grass root tips (see Dix & Webster, 1995).<br />

Other common species of Marasmius include<br />

M. androsaceus with thin, black, horsehair-like<br />

rhizomorphs and small basidiocarps arising<br />

from pine needles, dead heather, etc., and<br />

M. ramealis which forms clusters of basidiocarps<br />

on dead twigs and herbaceous stems. Marasmius<br />

rotula grows in similar situations. Its gills are not<br />

attached directly <strong>to</strong> the stem but <strong>to</strong> a cylindrical<br />

collar.<br />

Oudemansiella (10 spp.)<br />

This is a genus containing temperate and<br />

tropical species. Some species, e.g. O. mucida<br />

(porcelain fungus), have a membranous ring<br />

on the stem, but in others, e.g. O. radicata, a<br />

ring is lacking. Oudemansiella mucida is parasitic<br />

on beech (Fagus sylvatica) and forms white<br />

basidiocarps with slimy caps on the branches of<br />

infected trees. Oudemansiella radicata is so called<br />

because the base of the stipe is prolonged in<strong>to</strong> a<br />

tapering pseudorhiza which connects with<br />

buried woody branches extending for several<br />

cm beneath ground level (Fig. 19.18c; see also<br />

Buller, 1931).<br />

Flammulina (10 spp.)<br />

The best known species is F. velutipes, the velvet<br />

shank or winter fungus, which grows on deciduous<br />

tree stumps and branches and fruits in<br />

winter, forming golden yellow basidiocarps<br />

singly or in clusters (Plate 9d). Basidiocarps<br />

can survive being frozen and soon continue <strong>to</strong><br />

discharge spores on thawing. In the past, most<br />

collections of Flammulina were named F. velutipes,<br />

but it is now believed that several taxa had<br />

been included under this name (see Hughes et al.,<br />

1999). The Latin and trivial names refer <strong>to</strong> the<br />

brown velvety hairs at the base of the stipe.<br />

The distribution of the fungus is throughout<br />

temperate regions of the Northern Hemisphere.<br />

An anamorph is produced as chains of dry<br />

arthroconidia which have already been described<br />

(see Fig. 18.14a). The basidiocarps of F. velutipes<br />

are edible and the fungus is cultivated for food in<br />

Japan as enoki-take (Chang & Miles, 2004). Spawn<br />

of the fungus develops on sawdust in plastic bags<br />

or on logs incubated at 21 24°C for 14 18 d and<br />

then subjected <strong>to</strong> a cold shock (4 10°C) for 3 5d<br />

<strong>to</strong> induce fruiting, which follows within 5 8dat<br />

a temperature of 10 16°C. Light is required for<br />

fruiting.<br />

Graviperception and gravitropism<br />

in Flammulina<br />

When normally erect basidiocarps of F. velutipes<br />

are displaced in<strong>to</strong> a horizontal position, they<br />

respond within about 3 h by bending upwards <strong>to</strong><br />

res<strong>to</strong>re their original orientation (Fig. 19.19).<br />

Experimental studies, some conducted in orbit<br />

in a space labora<strong>to</strong>ry, have helped elucidate the<br />

mechanism (Moore et al., 1996; Kern et al., 1998;<br />

Kern, 1999). Bending occurs by curvature of<br />

the stipe in a transitional zone (2 3 mm long)<br />

immediately beneath the cap, where the hyphae<br />

making up the stipe and cap intertwine.<br />

Although basidiocarps which develop on Earth<br />

are erect, those formed in space were randomly<br />

orientated, growing out in all directions from<br />

their substrate. Despite this, normal caps, basidia<br />

and spores were produced. This shows that<br />

gravitational force is needed for stipe orientation<br />

but not for other aspects of fruit body development.<br />

Two responses operate <strong>to</strong> control fruit<br />

body development, namely negative hydrotropism<br />

causing basidiocarps <strong>to</strong> grow away from<br />

their moist substrate in<strong>to</strong> drier air, and negative<br />

gravitropism. Bending is caused by greater<br />

enlargement of the cells making up the lower<br />

flank of the stipe as compared with those of the<br />

upper flank. The cells of the lower flank are more<br />

vacuolate than those of the upper. In green<br />

plants with gravitropic responses, graviperception<br />

is correlated with the sedimentation of<br />

denser cy<strong>to</strong>plasmic particles, sta<strong>to</strong>liths, which


EUAGARICS CLADE<br />

549<br />

Fig19.19 Gravitropism of basidiocarps in<br />

Flammulina velutipes growing on a branch of<br />

gorse (Ulex europaeus).The branch was turned<br />

clockwise by 90° 4 h earlier, and the upper<br />

stipe regions are undergoing curvature<br />

(arrows) <strong>to</strong> bring the pilei back in<strong>to</strong> the<br />

horizontal position.<br />

congregate and are more abundant in the lower<br />

sides of cells in horizontally displaced organs. In<br />

Flammulina, Monzer (1996) considered that only<br />

the nuclei, with a density of 1.22 g cm 3 , were<br />

valid candidates for gravity-related sedimentation<br />

<strong>to</strong> enable them <strong>to</strong> function as sta<strong>to</strong>liths.<br />

In the transition zone there may be up <strong>to</strong> 10<br />

nuclei per hyphal segment. These nuclei are<br />

enmeshed in filaments of F-actin, and Monzer<br />

(1995) has suggested that tension of the actin<br />

filaments associated with sedimentation of<br />

nuclei is transmitted <strong>to</strong> the plasma membrane<br />

and provides the trigger <strong>to</strong> initiate cellular<br />

changes involved in the gravitational response.<br />

Evidence supporting the involvement of actin<br />

filaments is that the gravitropic response<br />

is affected by the actin-depolymerizing drug<br />

cy<strong>to</strong>chalasin D, but not by treatment with<br />

microtubule-inhibiting drugs.<br />

Within 30 min of displacement of a basidiocarp<br />

from a vertical <strong>to</strong> a horizontal position,<br />

differences may be noted between the upper and<br />

lower flank cells. Microvesicles, most likely<br />

derived from the endoplasmic reticulum and<br />

Golgi cisternae, are more abundant in the lower<br />

flank cells. They fuse with the vacuoles, thereby<br />

contributing <strong>to</strong> the volume increase and turgordriven<br />

enlargement of these cells. Vesicles<br />

containing wall precursor materials and<br />

enzymes also develop, providing for enlargement<br />

and stretching of the walls of lower flank cells.<br />

There are some 1.2 million hyphae present in the<br />

cross-section of a Flammulina stipe. They remain<br />

strictly parallel <strong>to</strong> each other but do not show<br />

anas<strong>to</strong>mosis or direct contact with each other.<br />

It is believed that each separate hypha in the<br />

transition zone responds <strong>to</strong> the gravitational<br />

stimulus. Possibly a growth fac<strong>to</strong>r which inhibits<br />

the growth of upper flank cells is involved, but<br />

no such substance has yet been demonstrated or<br />

identified.<br />

Detailed studies of gravitropism have also<br />

been made using stipes of the coprophilous inkcap<br />

Coprinus cinereus (Kher et al., 1992). Although<br />

there are similarities <strong>to</strong> Flammulina, there are<br />

also differences. For example, the response<br />

time of Coprinus is much shorter than that<br />

of Flammulina, and curvature extends along the<br />

entire stipe rather than being restricted <strong>to</strong><br />

the transition zone. These differences reflect<br />

the varied growth conditions of the two fungi<br />

(Moore et al., 1996). Flammulina velutipes is<br />

lignicolous and its basidiocarps are relatively<br />

long-lived whilst C. cinereus is coprophilous and<br />

its basidiocarps are evanescent.<br />

Armillaria (42 spp.)<br />

Most species are root pathogens, especially of<br />

woody plants, and there is an extensive literature<br />

on their biology and pathogenicity (see Shaw &<br />

Kile, 1991; Holliday, 1998; Fox, 2000). In the past,<br />

most collections were identified as A. mellea<br />

(the ‘honey fungus’), but this is now regarded<br />

as an aggregate of about 5 10 species, A. mellea


550 HOMOBASIDIOMYCETES<br />

agg. (Fig. 19.18d). Pegler (2000) has given a key <strong>to</strong><br />

the European species. Basidiocarps may occur<br />

singly, but often grow in clumps of dozens or<br />

even hundreds on the stumps of dead trees. The<br />

fruit bodies are reported <strong>to</strong> cause gastro-intestinal<br />

upsets if eaten raw, but are edible if they are<br />

cooked and the cooking water is discarded. Most<br />

species are annulate (i.e. with a ring on the stem),<br />

but in A. tabescens a ring is lacking. Serious<br />

pathogens include A. luteobubalina and A. mellea<br />

sensu stric<strong>to</strong> on a very broad range of hosts, and<br />

A. os<strong>to</strong>yae on conifers (Larix, Pinus, Pseudotsuga,<br />

Picea) and birch (Betula). Armillaria cepistipes and<br />

A. lutea are mainly saprotrophic.<br />

In infected trees, sheets of white mycelium<br />

grow between the wood and bark, destroying the<br />

phloem and cambium. In plantations, infections<br />

arise when air-borne basidiospores colonize the<br />

cut ends of thinning stumps. Spread of infections<br />

is by root-<strong>to</strong>-root contact from diseased <strong>to</strong><br />

healthy trees and by means of bootlace-like<br />

rhizomorphs (see and Fig. 18.13b). In this way<br />

adjacent trees in plantations become infected,<br />

resulting in group dying. A considerable area of<br />

forest may be affected by Armillaria, with an<br />

individual clone of A. bulbosa shown <strong>to</strong> extend<br />

over several hectares (Smith et al., 1992). The<br />

mycelium of Armillaria continues <strong>to</strong> grow saprotrophically<br />

after the death of an infected tree,<br />

and the zone of infected wood within the tree<br />

may be surrounded by dark, melanized, pseudosclerotial<br />

plates. Rhizomorphs may survive there<br />

for up <strong>to</strong> 40 years and continue <strong>to</strong> support<br />

fruiting over several years. Infected wood is<br />

often bioluminescent. Survival of the mycelium<br />

and its protection by pseudosclerotial plates<br />

makes the control of diseases caused by<br />

Armillaria difficult and expensive, although<br />

fungicidal treatment has been attempted (West,<br />

2000). Alternative biological control measures<br />

using fungal antagonists such as Trichoderma spp.<br />

also hold promise. Integrated control based on a<br />

combination of fungicidal treatments and biological<br />

antagonists (fungi and nema<strong>to</strong>des) has<br />

had a limited measure of success (Raziq, 2000).<br />

Although A. mellea agg. is generally regarded<br />

as a pathogen, it forms endotrophic mycorrhiza<br />

with several genera of chlorophyllous<br />

orchids in the tropics, and also with the<br />

colourless (achlorophyllous) orchid Gastrodia<br />

elata. As in other orchids (see p. 597) the seedlings<br />

of Gastrodia only become established following<br />

infection by the haploid basidiomycete<br />

Rhizoc<strong>to</strong>nia. However, the Rhizoc<strong>to</strong>nia infection is<br />

only a primary phase of limited duration, and<br />

secondary infection by Armillaria is essential<br />

for successful further growth of the orchid<br />

pro<strong>to</strong>corm. The Armillaria mycelium within the<br />

pro<strong>to</strong>corm is connected <strong>to</strong> mycelium growing<br />

parasitically on adjacent trees (Smith & Read,<br />

1997).<br />

Crinipellis (75 spp.)<br />

Crinipellis perniciosa is the cause of a severe<br />

witches’ broom disease of cocoa (Theobroma<br />

cacao) especially in South America (Purdy &<br />

Schmidt, 1996; Griffith et al., 2003). Young<br />

shoots are induced <strong>to</strong> proliferate and become<br />

swollen (Fig. 19.18e). There is considerable<br />

reduction in the yield of cocoa pods, and crop<br />

losses of up <strong>to</strong> 80% have been reported. Crinipellis<br />

perniciosa is hemibiotrophic; its monokaryotic<br />

mycelium is biotrophic but its dikaryotic mycelium<br />

is necrotrophic. It forms groups of small<br />

(pileus diameter: 5 15 mm) crimson <strong>to</strong> pink<br />

basidiocarps on dead cocoa twigs (Fig. 19.18f).<br />

Infection is by air-borne basidiospores which<br />

germinate under conditions of high humidity<br />

(e.g. dew) on meristematic tissues such as<br />

flowering cushions, young leaves and shoots.<br />

Germ tubes penetrating s<strong>to</strong>mata or wounded<br />

tissue establish a monokaryotic mycelium which<br />

pervades the proliferating hypertrophied shoots.<br />

Infection is not systemic. The monokaryotic<br />

phase does not grow in agar culture, but<br />

growth can be induced in cocoa or pota<strong>to</strong><br />

callus tissue cultures. There are several biotypes<br />

of C. perniciosa infecting different host plants.<br />

The C (cacao) biotype is primarily homothallic.<br />

Infected host twigs and pods eventually become<br />

necrotic and death of host tissues is associated<br />

with a change in the nuclear condition of the<br />

mycelium <strong>to</strong> the dikaryotic state. Basidiocarps<br />

develop on the dikaryon in the dead twigs. The<br />

dikaryotic mycelium can be grown in culture,<br />

and fruit body development is also possible on<br />

agar or on a bran vermiculite medium covered<br />

with a peat-based casing soil.


EUAGARICS CLADE<br />

551<br />

Control of witches’ broom disease is difficult<br />

but crop losses can be reduced by sanitation,<br />

i.e. the regular removal of brooms and diseased<br />

pods. Other measures involve the use of<br />

Trichoderma isolates as agents of biological<br />

control. Some are antagonistic, but T. stromaticum<br />

is mycoparasitic on mycelium and basidiocarps<br />

(Samuels et al., 2000; Sanogo et al., 2002).<br />

Systemic fungicides are also used. Further,<br />

there is an ongoing search for hosts resistant <strong>to</strong><br />

the pathogen which could be used for breeding<br />

purposes. The origin of the cocoa pathogen is<br />

thought <strong>to</strong> be native (wild) species of Theobroma<br />

growing in the forests, but other C. perniciosa<br />

strains attack members of the Solanaceae, Bixa<br />

spp., lianas (climbers) and woody debris on the<br />

forest floor.<br />

A closely related fungus, C. roreri<br />

(Myceliophthora roreri), causes the damaging<br />

frosty pod rot disease of cacao. Infected pods<br />

are covered by white powdery masses of winddispersed<br />

spores, originally regarded as conidia.<br />

These spores are borne on a dikaryotic mycelium<br />

which can be grown in culture. Dikaryotic<br />

hyphae swell and branch <strong>to</strong> form sporophore<br />

initials. The nuclei in the dikaryon fuse and septa<br />

are laid down, resulting in chains of diploid cells<br />

which have been interpreted as the equivalent of<br />

probasidia. The diploid cells develop thick walls<br />

<strong>to</strong> become spores and their nuclei undergo<br />

meiosis, but division may be arrested at a<br />

binucleate state (after the first meiotic division)<br />

or proceed <strong>to</strong> the formation of nuclear tetrads.<br />

On germination, there are indications that a<br />

four-celled metabasidium may be produced, with<br />

sterigmata which function as infective hyphae.<br />

As in C. perniciosa, this monokaryophase is biotrophic<br />

and can only be grown on living cocoa<br />

tissue. The life cycle of C. roreri is thus considerably<br />

modified, and a recognizable mushroomlike<br />

basidiocarp stage probably does not occur<br />

(Evans et al., 2002, 2003; Griffith et al., 2003).<br />

19.4.10 Mycenaceae<br />

Mycena (about 150 spp.)<br />

This is a large polyphyletic genus of fungi, with<br />

most species clustering in phylogenetic analyses<br />

around the type-species, Mycena galericulata. This<br />

group has been called Mycenaceae by Moncalvo<br />

et al. (2002). Mycena spp. produce rather small,<br />

delicate basidiocarps which have long slender<br />

stipes and conical or bell-shaped caps. Fruit<br />

bodies may emerge singly or in clusters from<br />

wood, leaf litter and other debris such as twigs,<br />

pine cones and bracken petioles in woodland and<br />

pastures. Some species exude latex when the<br />

stipe is broken, e.g. M. sanguinolenta with blood<br />

red latex and M. galopus which produces a<br />

milk white exudate. Much is known about the<br />

autecology of M. galopus, which has a perennial<br />

mycelium when growing in coniferous litter and<br />

also fruits on a wide range of twiggy debris but<br />

does not grow in bulky wood masses, possibly<br />

because of restricted aeration there (Frankland,<br />

1984; Dix & Webster, 1995). It is capable of<br />

growing on most of the constituents of leaf litter<br />

and its ability <strong>to</strong> break down lignin and cellulose<br />

enables it <strong>to</strong> function as a typical white-rot decay<br />

fungus. Mycena galopus is a key decomposer of<br />

oak leaves and, within two years of leaf fall,<br />

its mycelium may be present on 80% of fallen<br />

leaves. This species is regarded as a secondary<br />

colonizer, growing on plant material which is<br />

already well-colonized by other fungi. In this<br />

context, it is worth mentioning that many<br />

Mycena spp. produce antifungal metabolites,<br />

including strobilurins, and these may aid in the<br />

displacement of other wood-rotting fungi.<br />

Despite this, moribund fruit bodies of various<br />

Mycena spp., including strobilurin producers,<br />

may be parasitized by the zygomycete Spinellus<br />

fusiger (Plate 3e).<br />

Basidiocarp development in M. stylobates<br />

growing on beech leaves has been described<br />

and well-illustrated by Walther et al. (2001).<br />

An irregular arrangement of interwoven<br />

hyphae within the leaf bursts through <strong>to</strong> form<br />

an ovoid structure at the surface, composed<br />

mainly of vertically arranged hyphae. Cells at the<br />

margin increase in diameter and enclose the<br />

early stages of the primordium entirely.<br />

Separation of this large-celled wrapping tissue<br />

from internal hyphae results in the formation of<br />

a ring-like groove at the base of the primordium<br />

and a layer of protective hyphae covering a<br />

central bulb. The cells of the outer layer of<br />

vertically arranged hyphae increase in diameter


552 HOMOBASIDIOMYCETES<br />

<strong>to</strong> form the stipe. Simultaneously, hyphae at the<br />

apex of the bulb form horizontal outgrowths,<br />

giving rise <strong>to</strong> the pileus. The development of the<br />

hymenophore starts with the formation of small<br />

alveoli on the lower surface of the pileus, near its<br />

margin.<br />

19.4.11 Tricholomataceae<br />

This family has not yet been clearly separated<br />

from related groups in recent phylogenetic<br />

analyses. Genera currently placed here include<br />

Tricholoma, Lepista, Cli<strong>to</strong>cybe, Termi<strong>to</strong>myces and<br />

Lyophyllum. A few Collybia spp. (C. tuberosa,<br />

C. cirrhata, C. cookei) are also grouped here,<br />

whereas most of the genus Collybia sensu la<strong>to</strong><br />

seems <strong>to</strong> belong <strong>to</strong> the Marasmiaceae (see p. 546).<br />

It is probable that many other genera included<br />

in the Tricholomataceae are also polyphyletic<br />

(Moncalvo et al., 2000, 2002).<br />

Cli<strong>to</strong>cybe<br />

This is a large genus with about 60 species in<br />

Britain. The basidiocarps are funnel-shaped with<br />

decurrent gills (Fig. 19.20a) and can be large, up<br />

<strong>to</strong> 20 cm in diameter in C. geotropa. Cli<strong>to</strong>cybe<br />

nebularis may form fairy rings many metres in<br />

diameter in deciduous and coniferous woods.<br />

The fruit bodies of several species are edible and<br />

those of C. odora have a fragrant, aniseed-like<br />

flavour. Cli<strong>to</strong>cybe spp. are non-mycorrhizal.<br />

Lepista (50 spp.)<br />

The basidiocarps of Lepista are generally known<br />

as ‘blewits’ and include several good edible<br />

species such as L. nuda and L. saeva. The gills<br />

are attached <strong>to</strong> the stem in a sinuate manner,<br />

i.e. with an S-shaped point of attachment<br />

(see Fig. 19.8). Lepista is distinguished from<br />

Tricholoma in having a pale pink spore print<br />

(white in Tricholoma). Lepista spp. are nonmycorrhizal,<br />

growing mainly among humus<br />

and leaf litter.<br />

Tricholoma (200 spp.)<br />

This genus of mostly ec<strong>to</strong>mycorrhizal species is<br />

mainly North-temperate in distribution in woodlands<br />

and pastures. The basidiocarps have sinuate<br />

gills and produce a white spore-print.<br />

The most important edible species is the matsutake<br />

mushroom, T. matsutake, which is highly<br />

valued, especially in the Far East. Unfortunately,<br />

basidiocarps still have <strong>to</strong> be collected from<br />

woodlands because all attempts at cultivating it<br />

for commercial production have failed (see Wang<br />

& Hall, 2004). However, Guerin-Languette et al.<br />

(2005) have succeeded in infecting mature pine<br />

roots with T. matsutake in the labora<strong>to</strong>ry, and<br />

this may have laid the foundations for an<br />

inoculation pro<strong>to</strong>col for forest trees in future.<br />

An edible European species is St George’s mushroom<br />

(T. gambosum or Calocybe gambosa) which<br />

grows in pastures, often under trees, and fruits<br />

in spring around St George’s Day (23 April).<br />

Tricholoma sulphureum is a common woodland<br />

species whose basidiocarps are readily recognized<br />

by their sulphureous colour and unpleasant<br />

smell.<br />

19.4.12 Laccaria<br />

The relatively small basidiocarps of Laccaria<br />

(25 spp.) are known as ‘deceivers’ because of the<br />

change of colour intensity of the cap surface<br />

between wet (intensely coloured) and dry (pale)<br />

conditions. Laccaria spp. are important ec<strong>to</strong>mycorrhizal<br />

associates of forest trees and are used<br />

for artificial inoculation of trees <strong>to</strong> be planted<br />

out in<strong>to</strong> challenging situations. They are usually<br />

among the early-stage colonizers of tree<br />

roots, being replaced by others as the succession<br />

proceeds (Last et al., 1983, 1987; Kropp & Mueller,<br />

1999). Among the best known species are<br />

L. laccata, growing with Pinus and a range of<br />

other hosts, L. bicolor which grows with Abies,<br />

Pseudotsuga, Pinus and Picea, and L. amethystina<br />

(amethyst deceiver), an associate of oak (Quercus)<br />

and beech (Fagus). All species of Laccaria have<br />

edible basidiocarps. The relationships of Laccaria<br />

are poorly resolved, and family assignment is<br />

uncertain at present.<br />

Bertaux et al. (2003) have reported the<br />

presence of a bacterium (Paenibacillus sp.) within<br />

the hyphae of L. bicolor. Bacterial cells were<br />

visualized by fluorescence microscopy, and<br />

the escape of these bacteria in liquid culture<br />

conditions was shown <strong>to</strong> account for the sporadic<br />

bacterial ‘contamination’ of fermenter


EUAGARICS CLADE<br />

553<br />

Fig19.20 Basidiocarps in the euagarics clade (4). (a) Cli<strong>to</strong>cybe nebularis.(b)Stropharia semiglobata.(c)Psilocybe semilanceata,the<br />

magic mushroom. (d) Hypholoma sublateritium.(e)Cortinarius purpureus. Note the fibrillose cortina on the stipe of the young<br />

specimens (centre and right).<br />

cultures. Intrahyphal bacteria appear <strong>to</strong> be<br />

rare in higher fungi, but they have been reported<br />

in certain Zygomycota, notably Geosiphon (p. 221)<br />

and Rhizopus (p. 184).<br />

19.4.13 Strophariaceae<br />

Against all expectations, species included among<br />

the Strophariaceae do seem <strong>to</strong> group <strong>to</strong>gether,<br />

albeit with weak statistical support, in recent


554 HOMOBASIDIOMYCETES<br />

phylogenetic analyses (Moncalvo et al., 2002).<br />

Two groups of Psilocybe fall outside the core of<br />

the Strophariaceae, with the hallucinogenic<br />

species clustering separately from other<br />

Psilocybe spp.<br />

Stropharia<br />

Species of Stropharia fruit on soil, dung or wood.<br />

The developing gills are protected by a membranous<br />

veil which may persist as a solid annulus or<br />

a loose weft of fibrils. The spores are purplishbrown<br />

<strong>to</strong> black. Stropharia semiglobata, extremely<br />

common on several kinds of herbivore dung, has<br />

a viscid, yellowish, hemispherical cap and a ring<br />

of dark fibrils on the stem (Fig. 19.20b). Stropharia<br />

aeruginosa (verdigris agaric) has an attractive<br />

bluish-green viscid cap flecked with white<br />

scales and grows in woodland while S. aurantiaca,<br />

with bright orange caps, fruits on sawdust and<br />

on wood chippings.<br />

Psilocybe (c. 300 spp.)<br />

The basidiocarps of Psilocybe (Fig. 19.20c) are<br />

generally small and campanulate (bell-shaped),<br />

with attached gills and purple-brown spores.<br />

Some species have a fibrillose veil whilst others<br />

have a distinct annulus. The genus includes<br />

several species whose basidiocarps are hallucinogenic<br />

and are used as recreational drugs. Psilocybe<br />

mexicana (sacred mushroom, teonanácatl) has<br />

been used for centuries by Mexican Indians in<br />

religious ceremonies (Heim & Wasson, 1958),<br />

and P. cubensis (golden <strong>to</strong>p or giggle mushroom)<br />

is cultivated <strong>to</strong> obtain hallucinogenic basidiocarps.<br />

In Europe, P. semilanceata (magic mushroom,<br />

liberty cap) fruits in late summer and<br />

autumn on the ground in sheep pasture and<br />

well-manured grassland. It is saprotrophic, its<br />

mycelium being associated with decaying grass<br />

roots. Fruit bodies are picked and dried,<br />

although their sale is illegal in many European<br />

countries. The main hallucinogens of Psilocybe<br />

are the alkaloids psilocybin and psilocin<br />

(Figs. 19.15f,g), which are N-methylated tryptamines.<br />

They operate on sero<strong>to</strong>nergic systems of<br />

the brain and are similar in their effects <strong>to</strong><br />

mescaline and LSD (Lincoff & Mitchel, 1977;<br />

Bresinsky & Besl, 1990). Psilocin is less stable<br />

than psilocybin, and when oxidized shows a blue<br />

discoloration. This may be the reason why<br />

the flesh of many of the hallucinogenic Psilocybe<br />

spp. turns blue when bruised or broken.<br />

A readable account of the discovery of LSD<br />

and the elucidation of hallucinogenic principles<br />

in magic mushrooms has been written by<br />

the discoverer of both, Albert Hofmann. English<br />

translations of the original German text<br />

(Hofmann, 1979) are readily available.<br />

Hypholoma<br />

The best-known species are H. fasciculare (sulphur<br />

tuft) and H. sublateritium (Fig. 19.20d), both<br />

growing on wood. They are sometimes classified<br />

in the genus Naema<strong>to</strong>loma. Hypholoma fasciculare<br />

is very common, forming clusters of yellow<br />

basidiocarps on many kinds of deciduous and<br />

coniferous tree stumps, whilst H. sublateritium<br />

fruits on deciduous tree stumps and has brickred<br />

caps. There is a cot<strong>to</strong>ny veil on the stem and<br />

the spores are purplish-brown. Hypholoma fasciculare<br />

basidiocarps are inedible and bitter <strong>to</strong><br />

taste. They cause gastro-intestinal irritation and<br />

there are occasional reports of death following<br />

ingestion (Lincoff & Mitchel, 1977). Hypholoma<br />

fasciculare is highly competitive against other<br />

wood-rotting fungi. It is capable of extending<br />

from a woody food base through the soil by<br />

means of mycelial cords in search of further<br />

woody substrata. It can also utilize leaf litter. As<br />

a wood-decaying fungus its strategy is secondary<br />

resource capture, i.e. the displacement of<br />

other fungi which have earlier colonized wood,<br />

killing their mycelia. It therefore tends <strong>to</strong> fruit<br />

late in succession (Rayner & Boddy, 1988; Boddy,<br />

1993).<br />

Pholiota (c. 150 spp.)<br />

Most species of Pholiota grow on wood. Pholiota<br />

squarrosa is associated with soft, pale brown-rot<br />

(butt rot) of living trees of ash (Fraxinus), beech<br />

(Fagus) and poplar (Populus). Dense clusters of<br />

scaly brown basidiocarps are formed at the base<br />

of the trunk (Plate 9e). There is a prominent ring<br />

on the stem and the spores are smooth and<br />

rusty brown. It is best <strong>to</strong> avoid eating the<br />

basidiocarps of Pholiota spp. because serious<br />

ill-effects may follow, especially if they are<br />

consumed with alcohol. However, the nameko


BOLETOID CLADE<br />

555<br />

fungus (P. nameko) is cultivated commercially on<br />

sawdust-based substrate in Far Eastern countries<br />

for its edible basidiocarps (Chang & Miles, 2004).<br />

It is a moisture-loving fungus, growing in nature<br />

on dead trunks and stumps of deciduous<br />

trees at high altitudes in Japan and Taiwan. Its<br />

mating system is bipolar. Arthroconida develop<br />

on monokaryotic and on dikaryotic mycelia.<br />

Similarly, basidiocarps develop on monokaryons<br />

and on dikaryons.<br />

19.4.14 Cortinariaceae<br />

Cortinarius (c. 2000 spp.)<br />

Because of its large number of species, this genus<br />

provides one of the <strong>to</strong>ughest challenges <strong>to</strong> fungal<br />

taxonomists as well as field mycologists. In<br />

Britain alone, 230 species have been listed. The<br />

genus has been divided in<strong>to</strong> several subgenera<br />

(e.g. Cortinarius, Dermocybe, Leprocybe, Phlegmacium,<br />

Myxacium, Telamonia). Species in the subgenus<br />

Myxacium have slimy caps and stems derived<br />

from a glutinous universal veil, whilst in the<br />

subgenus Phlegmacium only the cap is sticky<br />

whereas the stem is dry. The genus is distributed<br />

in temperate regions of the Northern<br />

Hemisphere. All species are mycorrhizal with<br />

trees, so that they fruit in woodlands and<br />

woodland margins. The fruit bodies are small<br />

<strong>to</strong> large, buff, clay-coloured, orange-brown or<br />

sometimes very colourful, e.g. violet in C. violaceus<br />

or blood-red in C. sanguineus, C. purpureus and<br />

related species. Young fruit bodies are enveloped<br />

in a filamen<strong>to</strong>us or glutinous veil and developing<br />

gills are protected by a fibrillose cortina which<br />

may be evanescent or may persist, attached <strong>to</strong><br />

the stem (Fig. 19.20e). The spores are mostly<br />

warty, and the spore print is cinnamon <strong>to</strong> rustbrown.<br />

It is unwise <strong>to</strong> attempt <strong>to</strong> eat Cortinarius<br />

basidiocarps because the edibility of many<br />

species is unknown and some are deadly poisonous.<br />

The most no<strong>to</strong>rious among them is<br />

C. orellanus, which contains the bipyridyl <strong>to</strong>xin<br />

orellanine (Fig. 19.15h), the cause of delayed<br />

renal failure. Other <strong>to</strong>xins are cortinarins, which<br />

are cyclic peptides. The long delay of 2 20 days<br />

between ingestion of the mushroom and the<br />

onset of symp<strong>to</strong>ms (nausea, vomiting, diarrhoea,<br />

gastric upset and abdominal pain) are<br />

characteristic features of Cortinarius poisoning.<br />

Death may ensue 2 6 months later (Bresinsky &<br />

Besl, 1990; Michelot & Tebbett, 1990).<br />

19.5 Bole<strong>to</strong>id clade<br />

This clade includes not only the Boletales, but<br />

some other groups with basidiocarps which are<br />

dissimilar in appearance. Traditionally, the<br />

Boletales, typified by the genus Boletus, have<br />

included forms with fleshy, mushroom-like basidiocarps<br />

with tubular hymenophores. Later,<br />

based on morphological and chemical criteria,<br />

the concept was expanded <strong>to</strong> include gill-bearing<br />

forms such as Paxillus and Hygrophoropsis.<br />

Relationships were also suggested between<br />

poroid boletes and resupinate forms such as<br />

Coniophora or Serpula, and gasteroid genera such<br />

as Scleroderma or Rhizopogon. Molecular phylogenetic<br />

techniques have confirmed these relationships<br />

and a bole<strong>to</strong>id clade has been recognized <strong>to</strong><br />

embrace this wider concept (Bruns et al., 1998;<br />

Hibbett & Thorn, 2001). We shall study representatives<br />

of a gill-bearing group (Paxillaceae),<br />

poroid groups (Boletaceae, Suillaceae) and resupinate<br />

forms (Coniophoraceae).<br />

19.5.1 Paxillaceae<br />

Paxillus (15 spp.)<br />

Most species of Paxillus are ec<strong>to</strong>mycorrhizal<br />

(Wallander & Söderström, 1999), but P. atro<strong>to</strong>men<strong>to</strong>sus<br />

fruits on conifer stumps. The basidiocarps<br />

are soft and fleshy. A characteristic feature<br />

is that the hymenial tissue separates readily<br />

from the flesh of the cap. The roll-rim Paxillus<br />

involutus (Fig. 19.21a) has a brown, funnel-shaped<br />

cap with an inrolled margin, decurrent gills and<br />

brown spores. Upon bruising, an intense brown<br />

colour develops due <strong>to</strong> the accumulation of<br />

the pigment involutin (Fig. 19.22a), a member<br />

of the shikimic acid-derived pulvinic acid<br />

family typical of Boletales. Considering that it<br />

is a mycorrhizal species, P. involutus has an<br />

unusually wide host range, with 23 different<br />

tree species listed by Wallander and Söderström<br />

(1999). It is most commonly associated with birch<br />

(Betula) and oak (Quercus) in acid woodlands.


556 HOMOBASIDIOMYCETES<br />

Fig19.21 Basidiocarps in the bole<strong>to</strong>id clade. (a) Paxillus involutus. Note the inrolled cap margin and decurrent gills.The fruit<br />

bodies shown here have been attacked by the parasitic mould Sepedonium chrysospermum (white patches). (b) Boletus (Xerocomus)<br />

chrysenteron.(c)Boletus edulis, the cep or penny bun. (d) Suillusgrevillei, an ec<strong>to</strong>mycorrhizal associate of Larix.Notethering<br />

on the stem. (e,f) Serpula lacrymans, the dry rot fungus. (e) Beam supporting the roof of a church showing the typical<br />

cracking transverse <strong>to</strong> the grain of the wood which also shows shrinkage. (f) Resupinate fruit body on a ceiling showing the<br />

shallow pores.<br />

Spread through the soil <strong>to</strong> fresh young roots is<br />

by an effuse mycelium or by rhizomorphs.<br />

The fungus may survive in the soil by means of<br />

sclerotia. Some isolates are capable of saprotrophic<br />

growth and can fruit in the absence of a<br />

mycorrhizal host. Affinity with the Boletales has<br />

long been suspected because basidiocarps of P.<br />

involutus are commonly attacked by the bright<br />

yellow conidial state (Sepedonium chrysospermum)<br />

of the mycoparasitic ascomycete Apiocrea chrysosperma,<br />

which also attacks different boleti and<br />

gasteromycetes believed <strong>to</strong> be related <strong>to</strong> them<br />

(Fig. 19.21a; Plate 9h; p. 581).<br />

The basidiocarps of P. involutus, although traditionally<br />

considered edible, are, in fact, poisonous.<br />

There are two symp<strong>to</strong>ms. A heat-labile substance


BOLETOID CLADE<br />

557<br />

consumption. The <strong>to</strong>xic principle(s) appear <strong>to</strong> be<br />

unknown as yet.<br />

Fig19.22 Pulvinic acid-type pigments typical of members<br />

of the bole<strong>to</strong>id clade. (a) Involutin, a brown pigment produced<br />

by Paxillus involutus. (b) Variegatic acid, a yellow pigment.<br />

(c) The dark blue oxidation product of variegatic acid upon<br />

bruising the fruit bodies of Boletus erythropus and certain<br />

other Boletus spp. (d) Grevillin B produced by a range of<br />

Suillus spp.<br />

seems <strong>to</strong> be responsible for gastric upsets within<br />

a few hours of consuming raw or undercooked<br />

specimens, whereas haemolysis due <strong>to</strong> an<br />

allergic reaction can cause delayed liver<br />

failure and kidney damage only after repeated<br />

19.5.2 Boletaceae<br />

Boletus (300 spp.)<br />

Formerly comprising a broad range of porebearing<br />

fungi, the genus Boletus has been divided<br />

in<strong>to</strong> a number of groups, often now recognized<br />

as separate genera. Species of Boletus are ec<strong>to</strong>mycorrhizal.<br />

They have medium-sized, large or<br />

very large fleshy basidiocarps with a tubular<br />

hymenophore. The pores marking the openings<br />

of the hymenial tubes may be of the same<br />

yellowish-green colour as the tubes or may be<br />

coloured orange <strong>to</strong> blood-red as in B. erythropus<br />

(Plate 9f) and B. satanas. The pigments in Boletus<br />

are of the pulvinic acid group (Gill & Steglich,<br />

1987), with variegatic acid (Fig. 19.22b) being the<br />

most common. Within seconds of bruising or<br />

cutting the basidiocarps of certain species such<br />

as B. erythropus, their flesh and pores become<br />

discoloured blue or bluish-black due <strong>to</strong> the<br />

enzyme-mediated oxidation of variegatic acid<br />

(Fig. 19.22c) and xerocomic acid.<br />

One of the most common species, B. chrysenteron<br />

(Fig. 19.24b), sometimes classified in the<br />

separate genus Xerocomus, is easily recognized by<br />

the exposure of yellow or red flesh when the<br />

cap surface skin cracks. It is associated with<br />

broad-leaved trees. The stem of some species<br />

may be punctate, i.e. dotted with tiny warts (e.g.<br />

B. erythropus; Plate 9f), or veined (e.g. B. edulis).<br />

There are several species with delicious edible<br />

basidiocarps; amongst the best-known are B.<br />

edulis (cep or penny bun; Fig. 19.21c), B. badius<br />

(bay bolete; Plate 9h) and B. appendiculatus.<br />

Although it is possible <strong>to</strong> grow mycelium of B.<br />

edulis in pure culture, it is as yet impossible <strong>to</strong><br />

induce it <strong>to</strong> form basidiocarps, and therefore this<br />

much-prized edible species continues <strong>to</strong> be<br />

collected in forests (Wang & Hall, 2004). Some<br />

Boletus spp. are safe <strong>to</strong> eat only after cooking,<br />

e.g. B. luridus and B. erythropus, whereas others<br />

have poisonous basidiocarps, notably B. satanas<br />

(devil’s bolete) and B. satanoides.<br />

Boletus parasiticus forms fruit bodies attached<br />

<strong>to</strong> the basidiocarps of the earth ball Scleroderma.<br />

Doubts have been expressed as <strong>to</strong> whether this


558 HOMOBASIDIOMYCETES<br />

fungus is actually parasitic, and quite possibly it<br />

is merely stimulated <strong>to</strong> fruit by the presence of<br />

the earth ball.<br />

Leccinum (75 spp.)<br />

Like Boletus spp., the genus Leccinum has large<br />

fleshy basidiocarps. A characteristic feature is<br />

that the stem is covered with scales composed of<br />

cystidia. Leccinum scabrum (brown birch bolete)<br />

and L. versipelle (orange birch bolete) are common<br />

ec<strong>to</strong>mycorrhizal associates of birch (Betula).<br />

Basidiocarps of Leccinum are generally edible.<br />

19.5.3 Suillaceae<br />

Suillus (90 100 spp.)<br />

Suillus species have medium-sized fleshy basidiocarps<br />

forming ec<strong>to</strong>mycorrhizal associations with<br />

conifers. Their fruit bodies are usually of yellowish<br />

colours due <strong>to</strong> the abundance of pigments<br />

derived from the shikimic acid pathway, especially<br />

grevillins (Fig. 19.22d; Besl & Bresinsky,<br />

1997). The cap may be dry and scaly but is more<br />

usually viscid or slimy. There are species with a<br />

ring on the stem (Fig. 19.21d), e.g. S. grevillei<br />

(larch bolete) and S. luteus (slippery jack), whilst<br />

in many others (e.g. S. bovinus, S. granulatus) there<br />

is no ring (Plate 9g). The spores are smooth<br />

and elongate, and pale brown <strong>to</strong> brown. The<br />

presence of bundles of cystidia in the hymenium<br />

is a feature which distinguishes this genus<br />

from other boletes, and the classification in<strong>to</strong> a<br />

separate family is supported by phylogenetic<br />

studies (e.g. Grubisha et al., 2001). Also included<br />

in the Suillaceae are the gasteromycete genus<br />

Rhizopogon (see p. 581) and the gill-bearing<br />

Gomphidius.<br />

The distribution of species mirrors that of<br />

their mycorrhizal hosts, coniferous trees which<br />

are largely confined <strong>to</strong> North-temperate regions,<br />

only extending <strong>to</strong> other areas where introduced.<br />

There is a fairly high degree of host specificity.<br />

For example, in nature S. grevillei (Fig. 19.21d) is<br />

almost exclusively associated with Larix spp.<br />

(larch) whilst S. bovinus and S. luteus are associated<br />

with Pinus spp. This ecological specificity is<br />

probably associated with the effects of competition<br />

because in the labora<strong>to</strong>ry, under aseptic<br />

conditions, a wider range of hosts has been<br />

infected experimentally. Many species of<br />

Suillus have edible basidiocarps and some areas<br />

of planted pine (P. radiata) forests in South<br />

America are devoted <strong>to</strong> production of S. luteus,<br />

with as<strong>to</strong>nishing annual productivity values of<br />

up <strong>to</strong> 1 t dry weight ha 1 reported (Dahlberg &<br />

Finlay, 1999). Suillus bovinus and S. variegatus are<br />

common in Swedish pine plantations. Whereas<br />

S. bovinus is an early-stage mycorrhizal fungus,<br />

S. variegatus is often found in older stands.<br />

In open communities dominated by seedling<br />

pines, S. bovinus first develops a large number of<br />

genets (genetically defined mycelial individuals)<br />

derived from basidiospores. Once established,<br />

a colony extends by mycelial spread and by<br />

rhizomorphs <strong>to</strong> nearby host roots. The extent of a<br />

genet can be estimated by evidence of somatic<br />

incompatibility between mycelial isolates made<br />

from basidiocarps. As the original genets expand,<br />

they compete with each other. The number of<br />

genets decreases, but as the genets increase in<br />

size they may fragment so that parts of the same<br />

genet may become separated. The maximum<br />

rate of mycelial extension has been estimated <strong>to</strong><br />

be about 20 cm per annum and the age of the<br />

largest genet <strong>to</strong> be about 75 years (Dahlberg &<br />

Stenlid, 1990, 1994; Dahlberg, 1997).<br />

19.5.4 Coniophoraceae<br />

This group includes wood-rotting fungi such<br />

as Serpula and Coniophora which form<br />

spreading, crust-like resupinate fructifications.<br />

Morphological characteristics suggested that the<br />

Coniophoraceae have an affinity with Boletales<br />

(Pegler, 1991), and Hibbett and Thorn (2001)<br />

have included them in their bolete clade. This is<br />

supported by the presence of pulvinic acids in<br />

both Serpula and Coniophora (Gill & Steglich,<br />

1987). Here we shall consider only Serpula.<br />

Serpula<br />

Serpula lacrymans (Figs. 19.21e,f) causes dry rot<br />

and is one of the most serious agents of timber<br />

decay in buildings. There is a very extensive<br />

literature (see Rayner & Boddy, 1988; Jennings &<br />

Bravery, 1991). Both hardwoods and softwoods<br />

are attacked but, because softwoods are more<br />

commonly used in building construction, it is on


BOLETOID CLADE<br />

559<br />

such timbers that the fungus is most frequently<br />

reported. There is evidence that S. lacrymans<br />

has become adapted <strong>to</strong> man-made habitats.<br />

In the wild, S. lacrymans has been collected<br />

on spruce logs in the Himalayas at an altitude<br />

of 8000 10 000 ft (Bagchee, 1954; Singh et al.,<br />

1993), and it has also been reported from<br />

Northern Europe, Northern California and<br />

Siberia. It can continue <strong>to</strong> grow vegetatively at<br />

2°C, and the optimum (23°C) and maximum<br />

(26°C) temperatures for growth are rather low.<br />

Strains of S. lacrymans associated with houses<br />

are considered <strong>to</strong> belong <strong>to</strong> S. lacrymans var.<br />

domesticus and may have been spread by human<br />

activities in recent times. The Himalayas are<br />

often mentioned as a likely centre of origin,<br />

although Kauserud et al. (2004) put forward an<br />

alternative hypothesis featuring North America.<br />

Either way, possible vehicles for the dispersal<br />

of S. lacrymans var. domesticus may have been<br />

wooden sailing ships. Ramsbot<strong>to</strong>m (1938),<br />

in his fascinating review of a wealth of original<br />

literature sources, concluded that wooden<br />

vessels were frequently infested by S. lacrymans,<br />

often being unsound even before being launched<br />

due <strong>to</strong> the careless use of non-seasoned<br />

timber combined with poor ventilation of the<br />

ship holds. Typical symp<strong>to</strong>ms were described by<br />

a Commission of Inquiry, reporting <strong>to</strong> King<br />

James I in 1609 on the state of battleships in<br />

the Royal Navy (taken from Ramsbot<strong>to</strong>m, 1938):<br />

In buylding and repaireing Shippes with greene<br />

Tymber, Planck and Trennels it is apparent both by<br />

demonstration <strong>to</strong> the Shippes danger and by heate of<br />

the Houlde meeting with the greenesse and sappines<br />

thereof doth immediately putrefie the same and<br />

drawes that Shippe <strong>to</strong> the Dock agayne for<br />

reparation within the space of six or seven yeares<br />

that would last twentie if it were seasoned as it<br />

ought and in all other partes of the world is<br />

accus<strong>to</strong>med. Adde hereun<strong>to</strong> experience at this day<br />

that many Shippes thus brought in <strong>to</strong> be repaired,<br />

subject <strong>to</strong> miscareinge upon employment, and<br />

besides they breed infection among the men that<br />

serve in them.<br />

Only wood with a moisture content above about<br />

20 25% of the oven-dry weight is susceptible <strong>to</strong><br />

attack by the fungus. Well-dried and seasoned<br />

timber has a moisture content of 15 18%, and<br />

in a properly ventilated house this soon falls <strong>to</strong><br />

12 14% or lower. If woodwork becomes wet<br />

through contact with the soil, damp masonry,<br />

faulty construction or inadequate ventilation,<br />

then infection from air-borne basidiospores is<br />

likely <strong>to</strong> follow. Basidiospores germinate in the<br />

presence of free water on moist wood surfaces.<br />

The mycelium within the wood develops chiefly<br />

at the expense of the cellulose; lignin is not<br />

attacked, and the type of decay is a brown-rot.<br />

Well-rotted timber is shrunken with transverse<br />

cracks and has a dry crumbly texture.<br />

Water produced by the breakdown of cellulose<br />

(sometimes termed the water of metabolism)<br />

may be sufficient for further growth even if the<br />

air humidity is lowered below the point at which<br />

new basidiospore infections could arise. Up <strong>to</strong><br />

55.6% of the cellulose consumed may be available<br />

as metabolic water (see Bravery, 1991). As in<br />

many brown-rot fungi (see p. 527), oxalic acid is<br />

released in<strong>to</strong> the environment, lowering the pH<br />

of the wood and mortar over which S. lacrymans<br />

is growing.<br />

The epithet lacrymans (weeping) refers <strong>to</strong> the<br />

beads of moisture sometimes found on decaying<br />

timber, at the tips of hyphae and on mycelial<br />

cords. Sheets of mycelium may extend over the<br />

timber and adjacent brickwork, and the fungus<br />

is also capable of spreading several metres by<br />

means of mycelial cords up <strong>to</strong> 5 mm in diameter.<br />

The internal hyphae of the mycelial cords<br />

are exceptionally wide (up <strong>to</strong> 60 mm) and are<br />

modified for rapid conduction, enabling water<br />

and nutrients <strong>to</strong> be transported (Nuss et al., 1991).<br />

Transport is by pressure-driven hydraulic flow<br />

(Jennings, 1987, 1991). Carbohydrate is transported<br />

mainly in the form of trehalose.<br />

The strands can penetrate mortar and s<strong>to</strong>nework<br />

between walls and can spread throughout a<br />

building provided that there is enough wood as<br />

a food base. Strands remaining after removal<br />

of affected timber may still be able <strong>to</strong> initiate<br />

fresh infections.<br />

Reproduction in Serpula lacrymans<br />

This fungus is heterothallic and tetrapolar.<br />

Globally there are probably no more than four<br />

A and five B alleles (Schmidt & Moreth-Kebernik,<br />

1991). Arthroconidia are formed on


560 HOMOBASIDIOMYCETES<br />

monokaryons but not on dikaryons. Basidiocarps<br />

develop from dikaryons as flat, fleshy resupinate<br />

structures (Nuss et al., 1991). They may grow<br />

undetected under floorboards and in roof spaces<br />

for long periods and may reach a diameter of<br />

1 2 m. The lower side is brown and corrugated<br />

in<strong>to</strong> shallow pores supporting the hymenium<br />

(Fig. 19.21f). The folding of the hymenophore is<br />

the result of continuous thickening of the<br />

hymenium. The construction is at first monomitic,<br />

but becomes dimitic with the development<br />

of skeletal hyphae. Sporulation is<br />

continuous and it has been estimated that a<br />

basidiocarp measuring 100 cm 2 can produce<br />

300 million spores h 1 . The immense numbers<br />

of rusty-brown basidiospores may form deposits<br />

visible <strong>to</strong> the naked eye on cobwebs, shelves, etc.<br />

In the basements of buildings containing fruit<br />

bodies, spore concentrations of around<br />

80 000 m 3 air have been detected (see Hegarty,<br />

1991), raising concern over respira<strong>to</strong>ry allergy<br />

which may already have been referred <strong>to</strong> in the<br />

report <strong>to</strong> King James I (see above).<br />

Control of dry rot<br />

The economic consequences of failing <strong>to</strong> eradicate<br />

and control dry rot in a building can be<br />

severe. Modern methods aim <strong>to</strong> render the<br />

indoor environment hostile <strong>to</strong> S. lacrymans by<br />

eliminating all routes by which water could gain<br />

access <strong>to</strong> timber, accompanied by constantly<br />

high ventilation (Bravery, 1991; Palfreyman &<br />

White, 2003). A more traditional method is the<br />

removal of all infected timber and surrounding<br />

sound wood, and the treatment of any timber<br />

left in place with a recommended and approved<br />

fungicide. Replacement timbers should be<br />

selected for their durability properties and<br />

can be treated with pressure-impregnated fungicides<br />

such as copper/chromium/arsenic preservatives.<br />

Plaster can be treated with zinc<br />

oxychloride. Given the low temperature maximum<br />

of S. lacrymans, it is also sometimes possible<br />

<strong>to</strong> encase an entire building with a tent and<br />

subject it <strong>to</strong> thermal treatment (50°C). Since<br />

S. lacrymans is susceptible <strong>to</strong> attack by Trichoderma<br />

spp., biological control has been proposed but<br />

not yet put in<strong>to</strong> practice (Palfreyman et al., 1995).<br />

The most important control measure, however, is<br />

proper construction <strong>to</strong> ensure that the moisture<br />

level of the timber remains below the point at<br />

which infection can be initiated.<br />

19.6 Polyporoid clade<br />

Included in this clade are certain members of<br />

the Polyporales (¼ Aphyllophorales), a group<br />

comprising hymenomycetes in which (with a<br />

few exceptions) the hymenium is not borne on<br />

the surface of gills. It included bracket fungi<br />

(polypores), <strong>to</strong>oth fungi, coralloid fungi and<br />

forms with flattened or crust-like basidiocarps.<br />

However, morphological, ana<strong>to</strong>mical and chemical<br />

studies have indicated that this was not a<br />

natural grouping, and molecular phylogenetic<br />

investigations have amply confirmed this view,<br />

with Hibbett and Thorn (2001) showing that the<br />

aphyllophoroid condition occurs in all eight<br />

clades of Homobasidiomycetes. Most aphyllophoralean<br />

fungi outside the polyporoid clade<br />

are now considered <strong>to</strong> belong <strong>to</strong> the russuloid<br />

clade (see Section 19.7). The taxonomic hierarchy<br />

within the polyporoid clade is <strong>to</strong>o tentative <strong>to</strong> be<br />

adopted at present, and therefore we have<br />

desisted from using family names here.<br />

Economically important wood-rotting bracket<br />

fungi are found in the polyporoid clade. As<br />

described in detail on pp. 519 522, two main<br />

types of wood decay caused by basidiomycetes<br />

have been recognized, namely brown-rot in<br />

which cellulose is destroyed whereas lignin is<br />

left essentially unchanged, and white-rot in<br />

which both lignin and cellulose are attacked.<br />

Both types of rot can be caused by members of<br />

the polyporoid clade, e.g. brown-rot by Pip<strong>to</strong>porus<br />

betulinus (Fig. 19.23d) and white-rot by Trametes<br />

versicolor (Plate 10a). However, both rots can also<br />

be caused by other groups of basidiomycetes.<br />

The mycelium of members of the polyporoid<br />

clade is often perennial in large tree trunks and<br />

may give rise <strong>to</strong> a fresh crop of basidiocarps<br />

annually. In some species, e.g. Fomes fomentarius<br />

or Ganoderma applanatum (Figs. 19.23b,c), the<br />

fruit body itself may be perennial and new<br />

layers of hymenial tubes develop annually on<br />

the lower side of the basidiocarp. Typically


POLYPOROID CLADE<br />

561<br />

Fig19.23 Basidiocarps in the polyporoid clade. (a) Ganoderma lucidum. Basidiocarps formed in culture from sawdust contained<br />

in plastic bags. (b,c) Ganoderma applanatum. (b) Two sporophores attached <strong>to</strong> a living beech tree. (c) Detached sporophore split<br />

vertically <strong>to</strong> show two layers of hymenial tubes. (d) Pip<strong>to</strong>porus betulinus on a dead birch trunk. (e) Polyporus squamosus, basidiocarp<br />

attached <strong>to</strong> a living sycamore tree. (f) Polyporus brumalis growing from a dead beech twig. (g) Sparassis crispa fruiting at the base of<br />

a living pine tree. a kindly provided byY.-J.Yao.<br />

polypore fruit bodies develop as fan-shaped<br />

brackets lacking stipes, but there are forms<br />

with lateral (eccentric) stipes, such as Polyporus<br />

squamosus (dryad’s saddle; Fig. 19.23e), or even<br />

with centrally stalked fruit bodies, e.g. the<br />

winter polypore Polyporus brumalis (Fig. 19.23f).<br />

When the fruit bodies of polypores and other<br />

wood-rotting fungi develop on the underside of


562 HOMOBASIDIOMYCETES<br />

logs, they may be appressed <strong>to</strong> the surface of the<br />

wood and are then described as resupinate.<br />

The hyphal construction (hyphal analysis)<br />

of polypore basidiocarps varies (see p. 517 519)<br />

and is useful in identification. In some,<br />

e.g. Bjerkandera adusta, construction is monomitic;<br />

the basidiocarps are composed entirely of<br />

generative hyphae. The basidiocarps of Laetiporus<br />

sulphureus (chicken of the woods; Plate 10b) are<br />

dimitic, whereas Trametes versicolor has trimitic<br />

basidiocarps. The distinction between the different<br />

kinds of construction is best appreciated<br />

by attempting <strong>to</strong> tear the fruit bodies of these<br />

fungi apart. Trametes versicolor basidiocarps tear<br />

with difficulty, in contrast <strong>to</strong> the cheese-like<br />

consistency of L. sulphureus. Various modifications<br />

<strong>to</strong> the different hyphal systems may<br />

occur with age. For example, in L. sulphureus<br />

the generative hyphae may become inflated.<br />

In Polyporus squamosus the binding hyphae arise<br />

relatively late following inflation of the generative<br />

hyphae, converting the sappy flesh of the<br />

fully grown fruit body <strong>to</strong> a drier and firmer<br />

texture. In Pip<strong>to</strong>porus betulinus, <strong>to</strong>o, binding<br />

hyphae arise very late but ultimately replace<br />

the generative hyphae. The dissepiments (tissues<br />

between the pores) show a different construction,<br />

being dimitic with skeletal hyphae.<br />

The polyporoid clade is a large group,<br />

probably containing about 70 genera and over<br />

600 species. There is a very extensive literature.<br />

Notes on interesting features of some common<br />

polypores are given below.<br />

Trametes<br />

Trametes (Coriolus) versicolor (Plate 10a), colloquially<br />

called ‘turkey tail’, is a common saprotroph<br />

on various hardwood stumps and logs, causing<br />

white-rot. Both the mycelium and the fruit<br />

bodies are <strong>to</strong>lerant of desiccation. The annual<br />

fruit bodies have a zoned, multicoloured, velvety<br />

upper surface which readily absorbs rain. Details<br />

of the ana<strong>to</strong>my of the fruit body are shown in<br />

Fig. 19.24.<br />

Basidiocarps of T. versicolor may come in a<br />

range of colours and shapes, and if such<br />

variations in fruit body appearance occur on a<br />

single log, they can be traced <strong>to</strong> distinct columns<br />

of decayed wood when the log is serially<br />

sectioned. Dark brown ‘zone lines’ separate the<br />

columns. All isolations from within a column<br />

yield an identical dikaryon, and the zone lines<br />

mark intraspecific antagonism between distinct<br />

dikaryons (see Fig. 18.20a). Therefore, within a<br />

log of wood the fungus does not behave as<br />

a single ‘unit mycelium’ but as a series of<br />

discrete individuals (Rayner & Todd, 1977,<br />

1979). When monokaryons are inoculated experimentally<br />

in<strong>to</strong> logs in the field, they quickly<br />

become converted in<strong>to</strong> dikaryons by anas<strong>to</strong>mosis<br />

with compatible monokaryotic colonies derived<br />

from air-borne basidiospores (Williams et al.,<br />

1981). The individual dikaryotic colonies, once<br />

established, may persist and retain their integrity<br />

over several years, continuing <strong>to</strong> produce<br />

fresh crops of basidiocarps of the same genetic<br />

constitution each year. Antagonism, i.e. vegetative<br />

incompatibility between different dikaryons,<br />

can also be readily demonstrated in pure culture<br />

as shown in Fig. 18.20c. Monokaryotic mycelia<br />

form arthroconidia. This species is tetrapolar<br />

with multiple alleles.<br />

Because of its prolific production of<br />

peroxidase-type enzymes, T. versicolor is used<br />

industrially in such processes as the bioremediation<br />

of textile dyes or the wood preservative<br />

pentachlorophenol (PCP), and in delignification<br />

and decolorization of Kraft woodpulp. Some<br />

T. versicolor strains used in industry are thermo<strong>to</strong>lerant.<br />

In addition <strong>to</strong> enzymes, T. versicolor also<br />

produces polysaccharopeptides, and these are<br />

of commercial interest in anti-cancer therapy<br />

(reviewed by Cui & Chisti, 2003). A range of these<br />

substances are produced by different strains of<br />

the fungus. Their exact chemical composition<br />

is variable, with a branched sugar backbone<br />

consisting of b-(1,3) and a-(1,4) linkages and a<br />

protein content of about 30%. Polysaccharopeptides<br />

are extracted from mycelium grown in<br />

fermenters, purified, and administered orally.<br />

Although it is unclear how these large molecules<br />

can be taken up intact by the gut and how<br />

exactly they act <strong>to</strong> achieve the claimed results,<br />

several effects on the human body are suspected,<br />

with a general enhancement of the immune<br />

system being the most common. They are therefore<br />

considered useful as a complemention of<br />

other, more aggressive anti-cancer treatments.


POLYPOROID CLADE<br />

563<br />

Fig19.24 Trametes versicolor. (a) Vertical section through a basidiocarp. (b) Group of hyphae teased from the growing margin.Only<br />

generative and skeletal hyphae are present here. In the adult flesh, binding hyphae are also present (see Fig.19.3). (c) Transverse<br />

section across a pore showing the hymenium. (d) Longitudinal section of part of a dissepiment.The tissue contains only generative<br />

and skeletal hyphae. (b d) <strong>to</strong> same scale.<br />

Pip<strong>to</strong>porus<br />

Pip<strong>to</strong>porus betulinus basidiocarps are a common<br />

sight on dead and dying birch trees (Fig. 19.23d).<br />

The fungus is probably a wound parasite, its<br />

basidiospores entering and germinating where<br />

branches have broken off. Infected trees show<br />

a brown-rot of the heart wood which first<br />

undergoes cubical cracking and later disintegrates<br />

as a powder. Although the brown-rot<br />

indicates cellulose decay, there is also evidence


564 HOMOBASIDIOMYCETES<br />

of peroxidase activity which suggests limited<br />

lignin breakdown. The fungus is bipolar with<br />

about 30 alleles at the mating type locus.<br />

Multiple infections of single birch trunks are<br />

comparatively rare, but where they do occur,<br />

transverse sections of trunks with multiple<br />

infections show a characteristic pattern of<br />

black ‘zone lines’ marking the interface between<br />

antagonistic dikaryons (Adams et al., 1981). The<br />

dikaryons retain their integrity and persist over<br />

several years. A fresh crop of basidiocarps is<br />

produced annually on standing trees or on fallen<br />

trunks. The birch polypore was formerly known<br />

as the razor strop fungus in the days when the<br />

basidiocarps were used <strong>to</strong> hone ‘cut throat’<br />

razors.<br />

Laetiporus<br />

The poroid Laetiporus and Phaeolus, <strong>to</strong>gether with<br />

Sparassis which produces lobed, cauliflower-like<br />

fruit bodies (Fig. 19.23g), form one of the few<br />

well-resolved branches within the polyporoid<br />

clade (Wang et al., 2004). Biological features<br />

uniting these genera are that they cause wood<br />

decay of the brown-rot type often in living<br />

trees, and that their mating system is bipolar.<br />

The former species L. sulphureus sensu la<strong>to</strong> has now<br />

been subdivided in<strong>to</strong> several new taxa which<br />

show a certain degree of host specificity (Burdsall<br />

& Banik, 2001). Infection is typically through<br />

wounds of living trees, causing an intense<br />

brown-rot in the heartwood of standing trees.<br />

The mycelium is clearly long-lived, with fresh<br />

crops of basidiocarps (Plate 10b) produced<br />

annually for several years in the living tree,<br />

and later from the dead trunk. Laetiporus is<br />

considered <strong>to</strong> be one of the main causes for the<br />

hollowing of old oak trees in parks. Fruit trees<br />

(especially apple) are also affected. In the Alps,<br />

broad-leaved trees are attacked at altitudes<br />

below 3000 ft whereas coniferous species are<br />

infected higher up in the mountains. In addition<br />

<strong>to</strong> basidiospores, L. sulphureus also produces a<br />

chlamydosporic conidial state called Sporotrichum<br />

(see Fig. 18.16).<br />

Laetiporus sulphureus sensu la<strong>to</strong> has traditionally<br />

been considered an edible species, as its<br />

common name ‘chicken of the woods’ indicates.<br />

However, consumption has been associated with<br />

gastrointestinal upsets ( Jordan, 1995).<br />

Polyporus<br />

Polyporus squamosus (Fig. 19.23e) is a wound<br />

parasite of deciduous trees such as elm (Ulmus),<br />

beech (Fagus) and sycamore (Acer pseudoplatanus),<br />

producing an intensive white-rot. The mycelium<br />

persists on dead trunks, stumps and logs, and<br />

forms successive annual crops of basidiocarps<br />

during the early summer. The large, fan-shaped<br />

fruit bodies are creamy yellow, with brown<br />

scales. They are edible. Their texture is distinctly<br />

fleshy due <strong>to</strong> a dimitic structure with binding<br />

hyphae.<br />

Ganoderma<br />

There are more than 250 species of Ganoderma.<br />

They are white-rot fungi causing root and<br />

stem rots of hardwood and softwood hosts.<br />

A distinguishing feature is that the spore appears<br />

double-walled, with a dark-coloured inner layer<br />

bearing an ornamentation which pierces the<br />

hyaline outer one, so that the spore appears <strong>to</strong><br />

have a spiny surface (Fig. 19.25). Mims and<br />

Seabury (1989) have interpreted the basidiospore<br />

wall of Ganoderma lucidum as comprising three<br />

Fig19.25 Ganoderma applanatum. Basidiospore.The truncate<br />

portion is the apex of the spore. Spiny extensions of the<br />

darker inner spore wall penetrate the hyaline outer wall.


POLYPOROID CLADE<br />

565<br />

distinct components, an outermost primary<br />

wall, inter-wall pillars surrounded by electrontransparent<br />

regions, and an innermost secondary<br />

wall. Wall pillars, which initially develop<br />

immediately adjacent <strong>to</strong> the spore plasma<br />

membrane, eventually appear <strong>to</strong> fuse with both<br />

the primary and secondary walls.<br />

Ganoderma applanatum and G. adspersum are<br />

two species with similar basidiocarps which<br />

are confused with each other but can be<br />

distinguished by their basidiospores which are<br />

larger in G. adspersum. Both are wound parasites<br />

of deciduous trees, producing large (up <strong>to</strong> 1 m),<br />

perennial, brown, woody brackets (conks) which<br />

form a fresh layer of hymenial tubes annually<br />

(Figs. 19.23b,c). Recently formed hymenial tubes<br />

of G. applanatum are chocolate-brown in colour<br />

in contrast <strong>to</strong> the paler appearance of older<br />

layers formed earlier. The older layers are<br />

less dense and have a lower nitrogen content<br />

than the current year’s growth, suggesting<br />

that nitrogen is translocated from older,<br />

spore-depleted strata <strong>to</strong> the newly formed<br />

layers (Setliff, 1988). Nitrogen is often in<br />

limited supply in woody substrates, and its<br />

conservation is therefore important for wooddecaying<br />

fungi.<br />

Ganoderma applanatum is less common than<br />

G. adspersum and is usually found on old trunks<br />

of beech (Fagus sylvatica), where it causes a white<br />

heart rot. The hyphal structure of the fruit body<br />

is trimitic. A characteristic feature is that the<br />

skeletal hyphae are of two types, namely arboriform,<br />

i.e. showing an unbranched basal part<br />

with a branched tapering end, and aciculiform,<br />

i.e. unbranched and usually with a sharp tip<br />

(Hansen, 1958; Furtado, 1965). The hymenial<br />

tubes of G. applanatum may be up <strong>to</strong> 2 cm in<br />

length and about 0.1 mm in diameter, i.e. 200<br />

times as long as broad. The fall of the spores<br />

down this tube raises problems. The hard, rigid<br />

construction of the sporophore minimizes<br />

lateral disturbance <strong>to</strong> the vertical alignment of<br />

the tubes. Gregory (1957) has shown that the<br />

majority of the spores carry a positive electrostatic<br />

charge, but whether this charge has any<br />

relevance <strong>to</strong> the positioning of the spores during<br />

their fall by causing them <strong>to</strong> be repelled from the<br />

tube wall seems doubtful. It has been calculated<br />

that a large specimen may release as many as<br />

20 million spores min 1 during the 5 or 6 months<br />

from May <strong>to</strong> September (Buller, 1922). Spore<br />

discharge can continue even during periods of<br />

drought, doubtless associated with uptake of<br />

water from the tree host (Ingold, 1954b, 1957).<br />

In addition <strong>to</strong> wind dispersal, basidiospores of<br />

G. applanatum have been shown <strong>to</strong> be dispersed<br />

by specialized mycophagous flies visiting the<br />

basidiocarps (Tuno, 1999). Basidiocarps of<br />

G. applanatum are parasitized by larvae of the<br />

mycophagous fly Agathomyia wankowiczii. The<br />

trama is stimulated <strong>to</strong> proliferate in<strong>to</strong> conical<br />

or cylindrical gall-like outgrowths on the underside<br />

of the basidiocarp. When the larva has<br />

completed its development, it bores an exit hole<br />

through the tip of the gall and drops <strong>to</strong> the forest<br />

floor for pupation (Eisfelder & Herschel, 1966).<br />

Gall-forming insects rarely attack fungi, and<br />

even other Ganoderma spp. do not seem <strong>to</strong> be<br />

attacked by A. wankowiczii.<br />

Spores placed on media suitable for germination<br />

may take 6 12 months <strong>to</strong> develop germ<br />

tubes. Pairings between monosporous mycelia<br />

show that the fungus is tetrapolar, with multiple<br />

alleles. The large output of spores may be related<br />

<strong>to</strong> the low probability of compatible spores<br />

infecting the same tree trunk.<br />

As discussed above for Trametes, Ganoderma<br />

spp. have also been credited with having<br />

medicinal value against a wide range of<br />

ailments. Ganoderma lucidum (Fig. 19.23a), distinguished<br />

by its stalked, shiny (lacquered) basidiocarps,<br />

grows on the roots of deciduous trees.<br />

It is cultivated in China on sawdust in<br />

plastic bags for its basidiocarps from which<br />

polysaccharides and other pharmacologically<br />

active substances are extracted (Chang & Miles,<br />

2004).<br />

Phanerochaete<br />

There are about 100 species <strong>to</strong> this genus of<br />

saprotrophic fungi forming resupinate, crust-like<br />

basidiocarps with smooth, wrinkled or spiny<br />

but non-poroid hymenial surfaces. By far the<br />

best-known species is P. chrysosporium, which<br />

has become a ‘model’ organism for the examination<br />

of lignin-degrading enzymes (see Fig. 19.13)


566 HOMOBASIDIOMYCETES<br />

and their potential biotechnological applications.<br />

There is a conidial state called<br />

Sporotrichum pulverulentum (Burdsall, 1981). Little<br />

is known about the ecological role of P. chrysosporium<br />

because this species is rarely found in<br />

nature.<br />

19.7 Russuloid clade<br />

The russuloid clade is probably the most confusing<br />

group in the eight-clade system of Hibbett<br />

and Thorn (2001), containing the complete range<br />

of hymenophore types shown in Table 19.1 and<br />

Fig. 19.1. The phylogeny within this clade has<br />

been examined in detail by Larsson and Larsson<br />

(2003), who found that the main character uniting<br />

all members is the presence of gloeocystidia,<br />

i.e. cystidia filled with light-refractile (lipidrich)<br />

contents (Fig. 19.29b). In Stereum, these are<br />

modified as laticiferous hyphae (Fig. 19.27). Other<br />

characters commonly used <strong>to</strong> identify russuloid<br />

fungi are diagnostic only at lower taxonomic<br />

ranks, e.g. the ornamented walls of basidiospores<br />

and their starch-positive (amyloid) staining<br />

with Melzer’s iodine. We shall study representatives<br />

of agaricoid basidiocarps (Russulaceae),<br />

polyporoid fruit bodies (Heterobasidion), the<br />

spine-bearing Auriscalpium and corticioid forms<br />

(Stereaceae). The family arrangement is still<br />

tentative, and we give family names only where<br />

these have a fair chance of survival in future<br />

classification schemes.<br />

19.7.1 Russulaceae<br />

There are two important genera, Lactarius and<br />

Russula, and they share many features such<br />

as being ec<strong>to</strong>mycorrhizal species, and having<br />

a characteristically brittle texture <strong>to</strong> the gills<br />

and general fruit bodies due <strong>to</strong> the presence<br />

of sphaerocysts embedded in the tissue (see<br />

Fig. 19.4). Basidiospores are typically amyloid<br />

and ornamented with warts or ridges.<br />

Russula (c. 750 spp.)<br />

The genus Russula presents a similar taxonomic<br />

nightmare <strong>to</strong> Cortinarius, and it is unlikely that<br />

this genus is monophyletic. Species are widely<br />

distributed and ec<strong>to</strong>mycorrhizal. The basidiocarps<br />

(Fig. 19.26a) are moderate <strong>to</strong> large, often<br />

with a brightly coloured upper cap surface<br />

(white, yellow, green, red, purple or black). The<br />

gills are straight, arranged in a crowded but<br />

regular pattern, and vary from white <strong>to</strong> strawcoloured.<br />

The spores are ornamented with a<br />

network of branched ridges or plates. Some<br />

common species are R. ochroleuca (edible),<br />

R. fellea (bitter <strong>to</strong> the taste and inedible) and the<br />

sickener, R. emetica (poisonous). Russula ochroleuca<br />

fruits under coniferous and deciduous trees,<br />

R. fellea under beech (Fagus sylvatica) and R. emetica<br />

under pines.<br />

Lactarius (c. 400 spp.)<br />

The common name for Lactarius is milk cap,<br />

referring <strong>to</strong> the milky juice which exudes when<br />

the flesh of the cap is broken. The juice varies in<br />

colour from white <strong>to</strong> yellow, orange or violet<br />

and may change after exposure <strong>to</strong> the air. For<br />

example, the juice of L. deliciosus (saffron milk<br />

cap; Plate 10c) is carrot-coloured at first, but<br />

turns bright green upon prolonged exposure <strong>to</strong><br />

air. The juice also varies in taste, being mild <strong>to</strong><br />

faintly bitter in L. quietus, hot and acrid in<br />

L. pyrogalus, or first mild and then acrid in L.<br />

rufus. The juice is contained within broad<br />

laticiferous hyphae (see Fig. 19.4). As in Russula,<br />

the flesh also contains clusters of sphaerocysts.<br />

The gills are generally decurrent and the spores<br />

are ornamented like those of Russula. Some<br />

species of Lactarius have a narrow range of<br />

mycorrhizal partners, e.g. L. <strong>to</strong>rminosus (woolly<br />

milk cap) and L. turpis (ugly milk cap) are<br />

associated with birch (Betula), and L. deliciosus<br />

and L. deterrimus with Pinus and Picea. There<br />

are several species with edible basidiocarps,<br />

but some are poisonous or their edibility is<br />

unknown. Fruit bodies of L. deliciosus are especially<br />

valued, and it is possible <strong>to</strong> produce them<br />

in plantations of artificially inoculated<br />

host trees.<br />

19.7.2 Bondarzewiaceae<br />

Heterobasidion, a seemingly typical polyporoid<br />

wood-degrading bracket fungus, has been<br />

placed in the russuloid clade in several phylogenetic<br />

analyses. Some authors have assigned it <strong>to</strong>


RUSSULOID CLADE<br />

567<br />

Fig19.26 Basidiocarps in the russuloid clade. (a) Russula atropurpurea.(b)Heterobasidion annosum, the cause of butt rot of conifers.<br />

(c) Stereum hirsutum basidiocarps growing on an oak branch viewed from above. Note the white-rot at the broken end of the<br />

wood. (d) As (c) but viewed from below <strong>to</strong> show the smooth hymenium. (e) Auriscalpium vulgare basidiocarps growing from a<br />

buried pine cone.The hymenium is borne on spines on the lower side of the cap.<br />

the family Bondarzewiaceae (Larsson & Larsson,<br />

2003).<br />

Biology of Heterobasidion<br />

Heterobasidion annosum (Fig. 19.26b) is the cause of<br />

heart rot or butt rot of managed conifer<br />

plantations, and occasionally of deciduous trees<br />

such as birch in North-temperate regions<br />

(Korhonen & Stenlid, 1998; Asiegbu et al., 2005).<br />

It is the most important pathogen of conifers<br />

in these areas. The disease is especially common<br />

on alkaline soils. The fungus is a necrotrophic<br />

parasite whose mycelium extends downwards<br />

in<strong>to</strong> the root system and upwards in<strong>to</strong> the stem,<br />

causing a stringy heart rot which damages the<br />

strength and affects the saleability of the wood.<br />

Root decay is often followed by wind throw. The<br />

fungus may survive in the root system, especially<br />

if it is resinous, for over 60 years. In Europe, three<br />

somatic incompatibility groups have been distinguished,<br />

an ‘S’ group (S for spruce) confined <strong>to</strong><br />

Norway spruce (Picea abies), a ‘P’ group which


568 HOMOBASIDIOMYCETES<br />

Fig19.27 Stereum rugosum.<br />

Section of hymenium.The<br />

thick-walled hyphae with dense<br />

contents are sanguinolen<strong>to</strong>us<br />

hyphae which exude a red fluid<br />

when damaged, causing the<br />

hymenium <strong>to</strong> ‘bleed’.<br />

grows on pine, spruce, birch and other hosts,<br />

and an ‘F’ group growing on fir trees (Abies<br />

spp.) in Southern Europe. These groups, originally<br />

considered as infra-specific populations<br />

of H. annosum, have since been recognized as<br />

separate species (Niemelä & Korhonen, 1998).<br />

The name H. annosum is retained, but in<br />

a restricted sense for the ‘P’ group. The ‘S’ type<br />

is now named H. parviporum and the ‘F’ type<br />

H. abietinum. The first two groups also occur in<br />

North America.<br />

The fruit bodies are perennial, at first thin<br />

and leathery, later hard and woody (the epithet<br />

annosum means aged). They are formed at the<br />

base of dead trees, stumps or beneath fallen logs<br />

and are readily identified whilst still actively<br />

growing by their orange-brown colour with a<br />

white margin. They are dimitic with skeletals.<br />

Several successive layers of hymenial tubes are<br />

formed on old basidiocarps, and basidiospores<br />

are produced throughout the year. The mating<br />

system is unifac<strong>to</strong>rial (i.e. bipolar) with<br />

multiple alleles, probably over 100 (Korhonen &<br />

Stenlid, 1998). Basidiospores germinate <strong>to</strong><br />

form a homokaryotic primary mycelium with<br />

multinucleate segments. Anas<strong>to</strong>mosis with<br />

a compatible homokaryon results in the formation<br />

of a heterokaryotic secondary mycelium<br />

also with multinucleate segments and<br />

clamp connections at some, but not all, septa.<br />

The conidial state (Fig. 18.15) has been assigned<br />

<strong>to</strong> the anamorph genus Spiniger. Conidiophores<br />

can develop on primary and secondary mycelium<br />

and the conidia may be uninucleate or<br />

multinucleate.<br />

Freshly cut stump surfaces are hot-spots for<br />

infection by H. annosum, and the disease foci in<br />

plantations are often from thinning stumps<br />

which have become infected from air-borne<br />

spores. Insects are also possible vec<strong>to</strong>rs of<br />

conidia. Colonization of the roots of the stump<br />

is followed by spread <strong>to</strong> adjacent healthy trees by<br />

root-<strong>to</strong>-root contact and possibly by root-<strong>to</strong>-root<br />

graft because the fungus does not grow freely<br />

in the soil. Spread in this way may result in the<br />

development of genets (clones) many metres in<br />

diameter, in which several adjacent trees are<br />

infected by a secondary mycelium (heterokaryon)<br />

of identical genotype (Swedjemark & Stenlid,<br />

1993). The pathogenic effect of H. annosum<br />

may be correlated with the secretion of fungal<br />

<strong>to</strong>xins, of which several have been described,


RUSSULOID CLADE<br />

569<br />

predominantly benzofuran derivatives (Asiegbu<br />

et al., 1998).<br />

Control of Heterobasidion annosum<br />

Earlier methods of control were aimed at<br />

preventing stump infection by treatment with<br />

biocides such as creosote, sodium nitrite, urea,<br />

and boron-containing chemicals. The last two are<br />

still in use (Pratt et al., 1998; Asiegbu et al., 2005).<br />

An interesting alternative treatment, using<br />

biological control, is <strong>to</strong> inoculate freshly cut<br />

stumps with the conidia of the saprotrophic<br />

basidiomycete Phlebiopsis gigantea (¼ Peniophora<br />

gigantea; polyporoid clade) which competes<br />

with the parasite in the stumps and prevents<br />

its colonization, killing the cells of Heterobasidion<br />

by hyphal interference (Rishbeth, 1952, 1961;<br />

Ikediugwu et al., 1970). Conidial suspensions of<br />

P. gigantea are available commercially for use in<br />

biological control or, in conjunction with urea,<br />

in integrated control of stump infection in<br />

Picea plantations (Vasiliauskas et al., 2004, 2005).<br />

Trichoderma harzianum is another effective<br />

biological control agent (Holdenrieder & Greig,<br />

1998).<br />

19.7.3 Stereaceae<br />

In this family the fruit body is flattened,<br />

appressed or resupinate, with a smooth (untextured)<br />

hymenium on the lower surface.<br />

Construction may be monomitic or dimitic<br />

with skeletals. Basidiospores are commonly<br />

smooth-walled and amyloid. Most members are<br />

lignicolous and saprotrophic but some are<br />

pathogens of trees. Hibbett and Thorn (2001)<br />

have placed the family in the russuloid<br />

clade, and Larsson and Larsson (2003) found<br />

that there are several groups of Stereum-like fungi<br />

in the russuloid clade. Forms with Stereum-like<br />

basidiocarps also occur in several other clades of<br />

basidiomycetes (Yoon et al., 2003).<br />

Stereum sensu stric<strong>to</strong><br />

Basidiocarps of Stereum are common on decaying<br />

stumps and on attached and fallen tree branches.<br />

Stereum hirsutum (Figs. 19.26c,d) forms clusters<br />

of yellowish, fan-shaped leathery brackets with a<br />

hairy upper surface and a smooth hymenial<br />

surface on various angiospermous woody hosts,<br />

especially oak. It is a saprotrophic species<br />

and important as a cause of decay of sapwood<br />

of oak logs after felling. Stereum gausapatum is<br />

another common fungus on oak, which can grow<br />

parasitically on living trees, causing long narrow<br />

decay columns (‘pipe rot’) of the heartwood.<br />

When the hymenium is bruised it exudes a<br />

red latex. The same phenomenon is found in<br />

S. rugosum, common on coppice poles of hazel<br />

(Corylus avellana) and trunks of alder (Alnus<br />

glutinosa), and in S. sanguinolentum, a wound<br />

parasite or saprotroph on conifers. In most<br />

Stereum spp. there are specialized laticiferous<br />

(sanguinolen<strong>to</strong>us) hyphae which extend through<br />

the flesh in<strong>to</strong> the hymenium (Clémençon, 2004).<br />

These hyphae (Fig. 19.27) are thick-walled and are<br />

interpreted as modified skeletals homologous <strong>to</strong><br />

the gloeocystidia of other members of the<br />

russuloid clade.<br />

Mating in Stereum<br />

The mating behaviour of Stereum spp. presents<br />

some unusual features. Germinating basidiospores<br />

form a multinucleate primary mycelium,<br />

often with whorls (verticils) of clamp connections<br />

at the septa, a condition described as<br />

holocoenocytic (Boidin, 1971). This feature led<br />

<strong>to</strong> the erroneous conclusion that S. hirsutum<br />

is homothallic, but in fact it shows unifac<strong>to</strong>rial<br />

(i.e. bipolar) multi-allelic heterothallism (Coates<br />

et al., 1981). The mating type fac<strong>to</strong>r has been<br />

termed the ‘C’ fac<strong>to</strong>r. Compatible homokaryotic<br />

primary mycelia conjugate <strong>to</strong> form a heterokaryotic<br />

secondary mycelium which also produces<br />

whorled clamps. However, secondary mycelia can<br />

be distinguished by their yellowish pigmentation<br />

and often leathery surface, in contrast <strong>to</strong> the<br />

more delicate white primary mycelium. Some<br />

species of Stereum (S. hirsutum, S. sanguinolentum)<br />

include outcrossing and non-outcrossing populations<br />

(Ainsworth, 1987), whereas in others (e.g.<br />

S. gausapatum and S. rugosum) only outcrossing<br />

populations have been detected. In S. hirsutum,<br />

a feature of non-outcrossing homokaryons is<br />

their inability <strong>to</strong> accept non-self donor nuclei<br />

whilst they themselves can transfer nuclei <strong>to</strong><br />

outcrossing strains (Ainsworth & Rayner, 1989;<br />

Ainsworth et al., 1990). In the short term,


570 HOMOBASIDIOMYCETES<br />

non-outcrossing strains may have the selective<br />

advantage in being able <strong>to</strong> exploit a particular<br />

habitat. They are also resistant <strong>to</strong> potential<br />

takeover or conversion <strong>to</strong> unfit or unstable<br />

genomic combinations.<br />

Like many wood-rotting basidiomycetes,<br />

Stereum spp. show antagonistic reactions in the<br />

form of discoloured barrage zones where different<br />

strains confront each other, as seen in<br />

transverse sections of tree branches which<br />

contain more than one heterokaryon (Rayner &<br />

Boddy, 1988). Similar zones also develop on<br />

culture plates when dissimilar heterokaryons<br />

meet. Confrontations between heterokaryotic<br />

(secondary) and homokaryotic (primary) mycelia,<br />

equivalent <strong>to</strong> ‘di mon’ mating in Coprinus<br />

cinereus and Schizophyllum commune (see p. 508),<br />

have been studied in S. hirsutum in search of<br />

evidence of the ‘Buller phenomenon’ (Coates &<br />

Rayner, 1985a). Such confrontations may be fully<br />

compatible, where both types of nuclei from the<br />

heterokaryon are potentially competent in bringing<br />

about dikaryotization of the homokaryon, or<br />

hemi-compatible, where only one kind of<br />

nucleus is competent. The Buller phenomenon<br />

does occur in S. hirsutum, with either one or two<br />

new heterokaryons formed per interaction. The<br />

genotypes of the new heterokaryons can be<br />

classified as composite (accep<strong>to</strong>r homokaryon<br />

plus heterokaryon component), parental (identical<br />

with the donor heterokaryon) and novel<br />

(different from all three parental and composite<br />

combinations). There is preferential selection of<br />

a non-sib-related component of the donor heterokaryon,<br />

i.e. a nucleus from an unrelated strain,<br />

not derived from the same fruit body as the<br />

recipient homokaryon.<br />

The bow-tie reaction<br />

Because S. hirsutum is bipolar, 50% of sib-matings<br />

(i.e. matings between primary mycelia from<br />

basidiospores of the same basidiocarp) are<br />

compatible, whilst non-sib matings are uniformly<br />

compatible because of multiple alleles<br />

at the mating type locus. In about one-third of<br />

incompatible sib-related matings a distinctive<br />

pattern of mycelial interaction occurs in agar<br />

cultures termed the ‘bow-tie’ reaction (see<br />

Fig. 19.28a; Coates et al., 1981). A band of<br />

appressed sparse mycelium, shaped like a bowtie<br />

in being widest at the edges, develops<br />

between the two homokaryons, and is bounded<br />

by narrow regions with exude watery droplets.<br />

Hyphae within the bow-tie band often burst,<br />

have granular contents and produce abundant<br />

irregular lateral branches (Figs. 19.28b,c). It is<br />

believed that the bow-tie region of a culture<br />

is occupied by a weakly growing heterokaryon.<br />

The bow-tie region often expands and may<br />

replace the mycelium of one of the component<br />

monokaryons. In some cases, a darkened zone<br />

of mutual antagonism develops between them.<br />

The development of bow-ties is controlled by<br />

a multi-allelic genetic fac<strong>to</strong>r, the B-fac<strong>to</strong>r,<br />

unlinked <strong>to</strong> the mating type C-fac<strong>to</strong>r.<br />

Heterozygosity of the B-fac<strong>to</strong>r results in bowtie<br />

formation which is only visible between<br />

incompatible homokaryons. The bow-tie<br />

phenomenon, although discovered in S. hirsutum,<br />

is not a feature peculiar <strong>to</strong> this fungus. Similar<br />

behaviour has been found in mating type<br />

compatible pairings of some other basidiomycetes,<br />

e.g. S. gausapatum, Phanerochaete velutina,<br />

Mycena galopus and Coniophora puteana (Coates &<br />

Rayner, 1985b).<br />

Mating in Stereum sanguinolentum<br />

The mating behaviour of S. sanguinolentum is also<br />

unusual (Calderoni et al., 2003). Karyogamy,<br />

meiosis and post-meiotic mi<strong>to</strong>sis may occur in<br />

the four-spored basidia or spore primordia.<br />

Most basidiospores contain two nuclei and are<br />

heterokaryotic, producing a mycelium capable<br />

of fruiting. Stereum sanguinolentum therefore<br />

shows secondary homothallism or pseudohomothallism.<br />

Pairings between single-basidiospore<br />

isolates from the same basidiocarp (intrabasidiome<br />

pairings) are somatically compatible<br />

but virtually all pairings between isolates from<br />

different basidiocarps (inter-basidiome pairings)<br />

are somatically incompatible, with demarcation<br />

lines developing between them. This indicates<br />

that in S. sanguinolentum there are numerous<br />

vegetative incompatibility groups (Calderoni<br />

et al. 2003).


RUSSULOID CLADE<br />

571<br />

Fig19.28 The bow-tie reaction between incompatible monospore isolates of Stereum hirsutum. (a) General appearance during<br />

early development; w, watery exudation; b, bow-tie region. (b,c) Abnormal branching of hyphae in the bow-tie region.From<br />

Coates et al. (1981), with permission from Elsevier. Pho<strong>to</strong>graph kindly provided by A.D.M. Rayner.<br />

Chondrostereum purpureum<br />

There are several species superficially resembling<br />

Stereum but falling outside the core group<br />

currently placed in the russuloid clade. One<br />

such fungus is Chondrostereum purpureum<br />

(formerly known as Stereum purpureum), with<br />

purplish leathery bracket-like or resupinate<br />

basidiocarps (Plate 10d). Its exact placement in<br />

the Homobasidiomycetes remains <strong>to</strong> be determined.<br />

It is a wound pathogen of numerous<br />

genera of deciduous trees but especially of<br />

members of Rosaceae, including plum and<br />

cherry trees in which it causes ‘silver leaf’<br />

disease. The silver sheen on the leaves is due <strong>to</strong><br />

the separation of the epidermis from the<br />

mesophyll induced by secretions of <strong>to</strong>xic secondary<br />

metabolites (Strunz et al., 1997) and/or<br />

cellulase- and pectinase-type enzymes (Simpson<br />

et al., 2001) from mycelium in the branches<br />

below the leaves. Although C. purpureum is<br />

of considerable economic importance as a<br />

pathogen of fruit trees, it has been proposed as<br />

a biocontrol agent in forest situations where<br />

coniferous trees are <strong>to</strong> be grown and broadleaved<br />

trees, such as alder or birch, occur as<br />

weeds (Shamoun, 2000).<br />

Amylostereum<br />

There are several species of Amylostereum related<br />

<strong>to</strong> each other, but of uncertain placement within<br />

the Homobasidiomycetes (Slippers et al., 2003).<br />

They are necrotrophic wood-rotting wound<br />

pathogens of Pinus spp. and have entered a<br />

species-specific relationship with wood wasps<br />

of the genus Sirex, which distribute conidial<br />

(arthrosporic) inoculum in their internal glands<br />

and deposit it with mucilage surrounding<br />

their eggs, thereby providing the fungus with a<br />

suitable entry route in<strong>to</strong> host trees.<br />

Decomposition of the wood by Amylostereum<br />

spp. facilitates feeding by the Sirex larvae,<br />

which take up inoculum prior <strong>to</strong> pupation


572 HOMOBASIDIOMYCETES<br />

(Talbot, 1977). Severe infections of pine plantations<br />

are thought <strong>to</strong> be due <strong>to</strong> a synergism<br />

between the fungus and its insect vec<strong>to</strong>r,<br />

and are particularly common in the Southern<br />

Hemisphere.<br />

19.7.4 Auriscalpium<br />

The fruit bodies of Auriscalpium have a <strong>to</strong>othed<br />

hymenium, but molecular sequence studies have<br />

shown that the genus is related <strong>to</strong> the gillbearing<br />

Lentinellus and the clavarioid fungus<br />

Clavicorona (Pine et al., 1999), both included in<br />

the russuloid clade (Fig. 19.2). Auriscalpium<br />

vulgare, the earpick fungus, is distributed in the<br />

Northern Hemisphere. It grows on buried pine<br />

cones, forming stalked, one-sided, brown, hairy<br />

fruit bodies during autumn and winter<br />

(Fig. 19.26e). The hyphal construction is dimitic<br />

with skeletals. The hymenium is formed on<br />

vertical, finger-like downgrowths from the<br />

underside of the pileus. Interspersed amongst<br />

the basidia are irregularly enlarged, thin-walled<br />

hyphal tips with highly refractile contents,<br />

gloeocystidia (Fig. 19.29). Basidiocarps show<br />

rapid gravitropic readjustment <strong>to</strong> the vertical<br />

position if displaced laterally. The mating<br />

system of Auriscalpium is bifac<strong>to</strong>rial, i.e.<br />

tetrapolar (Petersen & Wu, 1992; Petersen &<br />

Cifuentes, 1994).<br />

19.8 Thelephoroid clade<br />

This includes the order Thelephorales, a small<br />

group of predominantly ec<strong>to</strong>mycorrhizal fungi<br />

with variable basidiocarps. The most important<br />

genus is Thelephora (c. 50 spp.). The earth fan<br />

T. terrestris (Plate 10e) produces clusters of fanshaped<br />

basidiocarps which are chocolate-brown<br />

in colour with a paler margin. They are<br />

often formed around the stem of young trees,<br />

seemingly ‘choking’ them. Basidiocarps of<br />

T. terrestris superficially resemble those<br />

of Stereum but are monomitic, composed of<br />

clamped generative hyphae only. The basidiospores<br />

are brown and warty. Thelephora terrestris<br />

fruits in association with coniferous trees<br />

growing on light sandy soils and heaths. It is<br />

Fig19.29 Auriscalpium vulgare.<br />

(a) Spine from the underside of the<br />

basidiocarp cap. (b) Portion of the<br />

hymenophore showing skeletal<br />

hyphae, gloeocystidia and basidia.


HYMENOCHAETOID CLADE<br />

573<br />

one of a group of early-stage ec<strong>to</strong>mycorrhizal<br />

associates of a variety of trees and also forms<br />

mycorrhiza with Arbutus menziesii, a member of<br />

the Ericaceae (Zak, 1976). According <strong>to</strong> Molina<br />

and Trappe (1984), T. terrestris represents the most<br />

abundant naturally present mycorrhizal fungus<br />

in bare-root nurseries.<br />

19.9 Hymenochae<strong>to</strong>id clade<br />

One feature that distinguishes the five<br />

Homobasidiomycete clades considered in the<br />

previous sections from the remaining three<br />

clades is the structure of the parenthesome, i.e.<br />

the membranous structure overarching the<br />

septal pore (see Fig. 18.10). In the five clades<br />

already described, the typical homobasidiomycete<br />

dolipore with a perforated parenthesome is<br />

found, whereas in the hymenochae<strong>to</strong>id, cantharelloid<br />

and gomphoid phalloid clades shown in<br />

Fig. 19.2, the parenthesome is generally imperforate<br />

(Hibbett & Thorn, 2001). Imperforate<br />

parenthesomes are also found in certain<br />

Heterobasidiomycetes, namely Dacrymycetales<br />

(Section 21.3) and Auriculariales (Section 21.4).<br />

This character has been discussed as being of<br />

phylogenetic significance, supporting the eightclade<br />

system shown in Fig. 19.2. Hibbett and<br />

Thorn (2001) have estimated that the hymenochae<strong>to</strong>id<br />

clade comprises about 630 spp.<br />

recruited from three families, namely the<br />

entire Hymenochaetaceae and parts of<br />

Corticiaceae and Polyporaceae.<br />

19.9.1 Hymenochaetaceae<br />

Most members of the hymenochae<strong>to</strong>id clade are<br />

wood-decomposing fungi, exemplified by the<br />

white-rots Inonotus and Phellinus which often<br />

fruit on old living trees and continue <strong>to</strong> do so<br />

after the death of the host. In general terms, the<br />

basidiocarps are monomitic and annual in<br />

Inonotus but dimitic, hard and perennial in<br />

Phellinus. Generative hyphae typically lack<br />

clamp connections in both genera. Fruit body<br />

morphology is variable, with resupinate and<br />

bracket-like forms most commonly produced;<br />

the hymenium is usually poroid. There are<br />

transitional forms between the genera Inonotus<br />

and Phellinus, and both have now been split up<br />

in<strong>to</strong> several smaller units (Wagner & Fischer,<br />

2001). Here we shall focus on Phellinus sensu la<strong>to</strong>.<br />

Phellinus sensu la<strong>to</strong> (c. 180 spp.)<br />

This genus is widespread and cosmopolitan,<br />

affecting both coniferous and broad-leaved<br />

trees. An interesting example is P. weirii, of<br />

which Hansen and Goheen (2000) have given<br />

a superb account. This species causes laminated<br />

root rot in native coniferous trees in the<br />

West Coast forests of North America, affecting<br />

especially Douglas fir (Pseudotsuga menziesii).<br />

Mycelium is able <strong>to</strong> penetrate intact roots and<br />

then spreads <strong>to</strong> adjacent trees by root-<strong>to</strong>-root<br />

contact. Trees of all ages are affected and<br />

are killed when extensive rot of the root<br />

system renders them unstable <strong>to</strong> s<strong>to</strong>rms, often<br />

several years or decades after initiation of<br />

infection. In this way, large genets of P. weirii<br />

slowly spread through the native forest,<br />

profoundly influencing the host population<br />

dynamics and shaping the forest landscape<br />

en route. Mycelium can also survive for decades<br />

in fallen logs. Phellinus weirii has a bipolar<br />

(unifac<strong>to</strong>rial) mating-system although, according<br />

<strong>to</strong> Hansen and Goheen (2000), basidiospores<br />

are not an important part of the life cycle which<br />

relies mainly on clonal spread.<br />

Phellinus noxius is the cause of root rot in<br />

numerous tree species in Central America and<br />

the Far East, Taiwan being particularly severely<br />

affected. Disease progression is unusually rapid<br />

for this group of pathogens, with infected trees<br />

sometimes dying within one growing season<br />

(Ann et al., 2002). This pathogen may remain<br />

viable in dead colonized roots for several years if<br />

the soil is relatively dry, and the flooding of<br />

affected areas, where practicable, is an efficient<br />

control method.<br />

There are also several species encountered in<br />

gardens, parks and forests in temperate climates,<br />

colonizing the trunks especially of mature and<br />

declining trees. Basidiocarps may be located<br />

several metres above ground. For example,<br />

P. igniarius (Plate 10f) forms basidiocarps predominantly<br />

on willow (Salix) and apple (Malus) trees,


574 HOMOBASIDIOMYCETES<br />

P. pomaceus on plum trees (Prunus), and P. robustus<br />

on oak (Quercus).<br />

Phellinus pomaceus has been extensively<br />

examined for the mechanism by which it<br />

produces the greenhouse gas chloromethane.<br />

This activity is indirectly associated with its<br />

lignin degradation system (see p. 527).<br />

19.10 Cantharelloid clade<br />

This is a further group of Homobasidiomycetes<br />

with a wide range of morphologically different<br />

basidiocarps. Species likely <strong>to</strong> be encountered<br />

during forays belong <strong>to</strong> the Cantharellaceae,<br />

Hydnaceae and Clavulinaceae. Pine et al. (1999)<br />

have provided evidence of the relationship<br />

between these groups. Hibbett and Thorn (2001)<br />

have suggested that Tulasnella might be included<br />

in this group, but since members of this genus<br />

produce heterobasidia, we regard them as<br />

belonging <strong>to</strong> the Heterobasidiomycetes (see<br />

Fig. 21.1b).<br />

19.10.1 Cantharellaceae<br />

The genera Cantharellus and Craterellus belong <strong>to</strong><br />

the most sought-after edible species. Cantharellus<br />

and related species are known as the chanterelles,<br />

with C. cibarius being the most readily<br />

recognized (Plate 10g). Another abundant,<br />

although less well-known, edible species is<br />

C. tubaeformis (Fig. 19.30a). The fruit bodies of<br />

most Cantharellus spp. appear in vibrant yellow or<br />

red colours due <strong>to</strong> carotenoid pigments, including<br />

canthaxanthin in the case of C. cinnabarinus.<br />

The fruit bodies are funnel-shaped, and the<br />

hymenium consists of shallow branching ridges<br />

which are strongly decurrent. The number of<br />

sterigmata and basidiospores on the basidium<br />

of C. cibarius seems <strong>to</strong> vary in the range of 4 7.<br />

In Craterellus cornucopioides, the horn of plenty<br />

or death’s trumpet, the hymenium is smooth,<br />

and the fruit bodies are trumpet-shaped with a<br />

hollow stipe.<br />

Members of this group are ec<strong>to</strong>mycorrhizal<br />

with coniferous (Picea, Pinus) and broad-leaved<br />

(Fagus) trees, and it has been possible <strong>to</strong><br />

grow C. cibarius on agar-based media and <strong>to</strong><br />

Fig19.30 Basidiocarps in the cantharelloid clade.<br />

(a) Cantharellus tubaeformis.(b)Hydnumrufescens, the hedgehog<br />

fungus. (c) Clavulina cristata.<br />

generate mycorrhizal associations under labora<strong>to</strong>ry<br />

conditions (Danell, 1994). There is hope<br />

that a method may ultimately be developed<br />

<strong>to</strong>wards commercial production of fruit


GOMPHOID PHALLOID CLADE<br />

575<br />

bodies under controlled conditions (Wang &<br />

Hall, 2004).<br />

19.10.2 Hydnaceae<br />

Hydnum<br />

The fruit bodies of the hedgehog fungi Hydnum<br />

repandum (see Fig. 19.1c) and H. rufescens<br />

(Fig. 19.30b) grow in deciduous and coniferous<br />

woodlands where they are ec<strong>to</strong>mycorrhizal. They<br />

are more or less mushroom-shaped, with a cap<br />

and a central or lateral stipe. Hydnum spp. are<br />

good <strong>to</strong> eat and H. repandum is often collected for<br />

sale in mainland Europe. The basidiocarp<br />

construction is monomitic, with generative<br />

hyphae which become inflated, giving the fruit<br />

body a fleshy texture. The hymenium covers the<br />

tapering spines which develop from the lower<br />

side of the cap.<br />

19.10.3 Clavulinaceae<br />

Although the spindle-shaped fruit bodies of<br />

Clavulina are morphologically very similar <strong>to</strong><br />

those of Clavaria, both genera have been placed<br />

in different clades. Clavulina spp. are saprotrophic<br />

fungi growing in the humus layer in<br />

forests and on lawns where they form coral- or<br />

spindle-shaped basidiocarps which are white,<br />

grey or pale yellow in colour. A common<br />

representative is C. cristata, which is a variable<br />

fungus with highly branched fructifications<br />

(Fig. 19.30c). A characteristic feature of the<br />

genus Clavulina is that the basidia are twospored,<br />

narrowly cylindrical, and often undergo<br />

septation after spore discharge. The hymenium<br />

thickens with age. The fruit body construction<br />

is monomitic, with clamped inflated hyphae<br />

(Fig. 19.31).<br />

19.11 Gomphoid phalloid clade<br />

There are about 350 species in this group which<br />

has given rise <strong>to</strong> the most fascinating array of<br />

gasteromycetes, including the cannonball<br />

fungus, earth stars and stinkhorns. These are<br />

described in detail in Section 20.4. There are<br />

also some genera of actively spore-discharging<br />

basidiomycetes in the gomphoid phalloid<br />

clade, especially with club-shaped and coralloid<br />

basidiocarps. These are grouped in the<br />

Clavariaceae which are briefly described below.<br />

19.11.1 Clavariaceae<br />

Hymenomycetes with smooth, branched or<br />

unbranched cylindrical or clavate fructifications<br />

were previously aggregated in<strong>to</strong> this family, but<br />

microscopical and molecular phylogenetic analysis<br />

show that such an arrangement groups<br />

<strong>to</strong>gether unrelated forms, and it is clear that<br />

the clavarioid type of fructification has evolved<br />

Fig19.31 Clavulinarugosa.Portionof<br />

the flesh and hymenium. Note the<br />

clamped hyphae and the narrow<br />

two-spored basidia.


576 HOMOBASIDIOMYCETES<br />

independently in several unrelated basidiomycete<br />

groups (Pine et al., 1999). For example,<br />

Clavulina is now included in the cantharelloid<br />

clade (p. 575). Corner (1950) has monographed<br />

the clavarioid fungi and Petersen (1973) has<br />

given keys <strong>to</strong> genera.<br />

Clavaria and its allies<br />

This is a large genus of pasture and woodland<br />

fungi with cylindrical or club-shaped branched<br />

or unbranched fructifications. The flesh of<br />

Clavaria is made up of thin-walled hyphae<br />

which lack clamp connections and may become<br />

inflated and develop secondary septa. The hymenium<br />

which covers the whole surface of the<br />

fruit body usually consists of four-spored basidia<br />

with or without basal clamps, bearing colourless<br />

spores. A typical species is C. vermicularis<br />

which fruits in grassland, forming tufts of<br />

whitish, unbranched spindle-shaped basidiocarps.<br />

Clavaria argillacea forms yellow clubshaped<br />

fructifications on moors, heaths and<br />

peat bogs. In North America fructifications are<br />

consistently associated with ericaceous plants,<br />

including cultivated blueberries (Vaccinium angustifolium<br />

and V. myrtilloides), Azalea and Erica<br />

(Englander & Hull, 1980). In Australia a similar<br />

species is associated with cultivated Azalea<br />

indica (Seviour et al., 1973). The close association<br />

suggests a mycorrhizal relationship, and<br />

although it has not been possible <strong>to</strong> synthesize<br />

mycorrhiza by inoculating aseptically grown<br />

ericaceous seedlings with mycelial cultures of<br />

Clavaria, evidence in support of a mycorrhizal<br />

partnership has been obtained by the demonstration<br />

of a two-way transfer of radioactive carbon<br />

and phosphorus between Clavaria and ericaceous<br />

plants, and by immunofluorescence studies in<br />

which hyphal coils were stained within root<br />

epidermal cells with conjugated antibodies<br />

raised against Clavaria basidiocarps (Mueller<br />

et al., 1986). It is therefore likely that ericoid<br />

mycorrhizae develop following root infection by<br />

basidiomycetes as well as by ascomycetes such as<br />

Hymenoscyphus ericae (see p. 442), and both groups<br />

of fungi may infect the same root.<br />

There are numerous other common representatives<br />

of the Clavariaceae. Clavariadelphus pistillaris<br />

(giant club) forms exceptionally large clubshaped<br />

fruit bodies (7 30 2 6 cm) in beech<br />

woods on chalk (see Fig. 19.1d). The construction<br />

is monomitic, with clamps at the septa. As the<br />

fruit body matures, the hymenium becomes<br />

thicker by the development of further layers of<br />

basidia. Some of the more richly branched fairy<br />

clubs are placed in the genus Ramaria, distinguished<br />

by <strong>to</strong>ugher flesh and pink, yellow or<br />

brown-coloured basidiospores, which are often<br />

rough. A particularly striking example is<br />

R. botrytis (Plate 10h). Most Ramaria spp. are<br />

ec<strong>to</strong>mycorrhizal (Nouhra et al., 2005); R. stricta is<br />

exceptional in growing on rotten wood.


20<br />

Homobasidiomycetes: gasteromycetes<br />

20.1 <strong>Introduction</strong><br />

Gasteromycetes are an unnatural assemblage of<br />

basidiomycetes sharing the common negative<br />

character that the basidiospores are not<br />

discharged violently from their basidia. Instead<br />

of the ballis<strong>to</strong>sporic basidiospores of other<br />

basidiomycetes which are asymmetric in side<br />

view (see Fig. 18.5), those of the gasteromycetes<br />

are usually symmetrically poised on their sterigmata<br />

or are sessile. Dring (1973) has termed<br />

such basidiospores statismospores. Commonly<br />

the basidia open in<strong>to</strong> cavities within a fruit body,<br />

and the basidiospores are released in<strong>to</strong> these<br />

cavities as the tissue between them breaks down<br />

or dries out. A recognizable fertile layer (hymenium)<br />

may be present or absent (see Reijnders,<br />

2000). The internal production of basidiospores<br />

has given the gasteromycetes their name (Gr.<br />

gaster ¼ s<strong>to</strong>mach). The gasteromycete fruit body<br />

is termed the gasterocarp, and the spore mass<br />

enclosed by the gasterocarp wall (peridium)<br />

is the gleba. Sometimes, as in Lycoperdon<br />

or Geastrum, the gasterocarp opens by a pore<br />

through which the spores escape, but in forms<br />

with subterranean (hypogeous) fruit bodies there<br />

is no special opening, and it is possible that the<br />

spores are dispersed by rodents and other<br />

burrowing animals (Colgan & Claridge, 2002).<br />

In Phallus and its allies, the basidiospores are<br />

exhibited in a sticky mass attractive <strong>to</strong> insects,<br />

whilst in Cyathus and Sphaerobolus the spores are<br />

enclosed in separate glebal masses or peridioles<br />

which are dispersed as units. In spite of these<br />

variations in gasterocarp morphology, the life<br />

cycles of most gasteromycetes follow the<br />

general Homobasidiomycete pattern outlined<br />

in Fig. 18.4. Most species for which details are<br />

known appear <strong>to</strong> be heterothallic, with a basidiospore<br />

germinating <strong>to</strong> give a monokaryotic<br />

primary mycelium. Following fusion of compatible<br />

primary mycelia, a dikaryotic secondary<br />

mycelium is established, and this produces<br />

gasterocarps in which karyogamy and meiosis<br />

occur, and haploid basidiospores are formed.<br />

Dikaryotic asexual propagules are also known in<br />

some gasteromycetes.<br />

Most members of the group are saprotrophic<br />

and grow on soil, rotting wood and other<br />

vegetation, or dung. Mycelial cords or rhizomorphs<br />

are often formed. Rhizopogon which<br />

produces hypogeous gasterocarps, and<br />

Scleroderma and Pisolithus with epigeous fruit<br />

bodies, are important ec<strong>to</strong>mycorrhizal associates<br />

of forest trees. There are also two genera of<br />

aquatic gasteromycetes. Nia vibrissa grows on<br />

driftwood in the sea, forming globose, yellowish<br />

gasterocarps a few millimetres in diameter. Its<br />

basidiospores bear 4 5 radiating appendages<br />

(Jones & Jones, 1993). Such appendages are a<br />

typical adaptation <strong>to</strong> the aquatic habitat (see<br />

Section 25.2). Limnoperdon forms small, floating<br />

fruit bodies in freshwater swamps and marshes<br />

(Escobar et al., 1976).<br />

Because of the conspicuous shape and appearance<br />

of their gasterocarps, gasteromycetes have<br />

long attracted the attention of mycologists, and<br />

several keys and descriptions are available,<br />

including the books by Miller and Miller (1988),


578 HOMOBASIDIOMYCETES: GASTEROMYCETES<br />

Ellis and Ellis (1990) and Pegler et al. (1995).<br />

Species with hypogeous gasterocarps (‘false<br />

truffles’) have been described by Pegler et al.<br />

(1993).<br />

20.2 Evolution and phylogeny of<br />

gasteromycetes<br />

In theory, the evolution of a gasteromycete<br />

from a hymenomycete ances<strong>to</strong>r requires only<br />

two morphogenetic changes, i.e. the production<br />

of a closed fruit body accompanied by the loss<br />

of the active spore discharge mechanism.<br />

The coincidence of these two features is shown<br />

by several examples of secotioid fruit bodies,<br />

i.e. basidiocarps in which the margin of the<br />

pileus fails <strong>to</strong> become detached from the stipe<br />

(Thiers, 1984). Watling (1971) described such<br />

aberrant development in the agaric Psilocybe<br />

merdaria growing in culture, where the failure<br />

of the fruit body <strong>to</strong> expose its hymenium<br />

coincided with morphological changes <strong>to</strong> the<br />

gills and basidia. Chiu et al. (1989) reported a<br />

similar case in Volvariella bombycina, and Hibbett<br />

et al. (1994) demonstrated that a naturally<br />

occurring recessive allele in a single gene is<br />

responsible for converting the lamellate fruit<br />

body of Lentinus tigrinus in<strong>to</strong> a secotioid one.<br />

Except for the L. tigrinus mutant in which an<br />

existing hymenium is belatedly overgrown by a<br />

veil (see Fig. 19.6b), most secotioid forms seem <strong>to</strong><br />

arise as a developmental defect causing incomplete<br />

differentiation of an agaric- or bolete-type<br />

basidiocarp (Thiers, 1984; Hibbett et al., 1997b).<br />

Several species pairs are known in nature in<br />

which a secotioid form is closely related <strong>to</strong><br />

a mushroom-type species, e.g. Montagnea and<br />

Podaxis (Fig. 20.1) related <strong>to</strong> Coprinus comatus<br />

(Fig. 19.14c), Gastroboletus related <strong>to</strong> Boletus,<br />

Gastrosuillus related <strong>to</strong> Suillus, Hydnangium related<br />

<strong>to</strong> Laccaria, orThaxterogaster related <strong>to</strong> Cortinarius<br />

(Thiers, 1984; Mueller & Pine, 1994; Hopple &<br />

Vilgalys, 1999). Such evolutionary trends<br />

<strong>to</strong>wards ‘gasteromycetation’ may be ongoing,<br />

and e.g. Gastrosuillus laricinus is thought <strong>to</strong><br />

have arisen from Suillus grevillei as recently as<br />

Fig 20.1 Fruit bodies of the secotioid<br />

mushroom Podaxis pistillaris, a close relative of<br />

the ink cap Coprinus comatus (see Fig.19.14c).<br />

(a) Young fruit body. (b) Mature disintegrating<br />

fruit body.Original pho<strong>to</strong>graphs kindly<br />

provided by A.E. Ashford.


EVOLUTION AND PHYLOGENY OF GASTEROMYCETES<br />

579<br />

70 years ago (Baura et al., 1992). In contrast, the<br />

most ancient fossil gasteromycete found so far,<br />

an earth star resembling Geastrum, dates back <strong>to</strong><br />

the Cretaceous period some 65 70 million years<br />

ago (Krassilov & Makulbekov, 2003).<br />

One selective environmental pressure <strong>to</strong>wards<br />

the secotioid and ultimately gasteromycete habit<br />

might be drought, since the very nature of the<br />

active basidiospore discharge mechanism by<br />

drop fusion (see p. 493) precludes its function at<br />

low humidity. It is perhaps no coincidence that<br />

secotioid fungi are particularly common in arid<br />

regions (Thiers, 1984). Secotioid fruit bodies are<br />

generally assumed <strong>to</strong> be the first step <strong>to</strong>wards<br />

typical gasteromycete forms such as earth balls,<br />

puffballs and false truffles (Reijnders, 2000).<br />

However, mycologists are still at a loss <strong>to</strong> explain<br />

how the fantastically complicated fruit bodies,<br />

e.g. of the stinkhorns or bird’s nest fungi, could<br />

have evolved from there.<br />

Given the ease with which secotioid fruit<br />

bodies can arise, it is hardly surprising that<br />

gasteromycetes have evolved several times independently<br />

from hymenomycete ances<strong>to</strong>rs, as<br />

indicated by numerous phylogenetic studies<br />

(see Fig. 19.2; Hibbett et al., 1997b; Hibbett &<br />

Thorn, 2001). In subsequent sections of this<br />

chapter we shall consider the three most<br />

important groupings which are as follows (see<br />

Table 20.1):<br />

1. Members of the euagarics clade (Section<br />

19.4). The puffballs (Lycoperdaceae) and bird’s<br />

nest fungi (Nidulariaceae) as well as a few<br />

smaller groups of gasteromycetes belong <strong>to</strong> this<br />

group. The Lycoperdaceae are close <strong>to</strong><br />

Macrolepiota (Krüger et al., 2001), whereas the<br />

Nidulariaceae cannot be placed accurately as yet<br />

but are likely <strong>to</strong> have arisen on a separate<br />

occasion. A further independent evolutionary<br />

event was that leading <strong>to</strong> the marine gasteromycete<br />

Nia vibrissa (Binder et al., 2001).<br />

2. Members of the bole<strong>to</strong>id clade (Section<br />

19.5). Several gasteromycetes have their origin<br />

in the bole<strong>to</strong>id clade (Binder & Bresinsky, 2002).<br />

The most important group is the family<br />

Sclerodermataceae, i.e. the earth balls and their<br />

relatives (Scleroderma, Pisolithus, Astraeus) which<br />

are closely related <strong>to</strong> Gyrodon. Another example<br />

is Rhizopogon which is close <strong>to</strong> Suillus. Like their<br />

actively spore-discharging relatives, these bole<strong>to</strong>id<br />

gasteromycetes are important ec<strong>to</strong>mycorrhizal<br />

associates of trees.<br />

3. The gomphoid phalloid clade. This<br />

group contains the coral fungi and similar<br />

basidiomycetes with exposed hymenia and<br />

active basidiospore discharge (Ramaria,<br />

Clavariadelphus, Gomphus; see p. 575), as well as<br />

several important groups of gasteromycetes,<br />

namely the earth stars (Geastrum spp.), the<br />

cannonball fungus (Sphaerobolus), and the stinkhorns<br />

and their allies. The phylogeny of this<br />

grouping has been discussed by Humpert et al.<br />

(2001).<br />

Although the artificial nature of the gasteromycetes<br />

as a taxonomic group has been known<br />

or suspected for many decades, it still comes as a<br />

shock <strong>to</strong> most mycologists <strong>to</strong> realize just how<br />

strongly convergent the evolution of these fungi<br />

has been. For instance, the implications from the<br />

results of phylogenetic studies (see Table 20.1)<br />

are that the earth stars (Geastrum spp.) have<br />

arisen independently of the barometer earth star<br />

(Astraeus), that the raindrop-mediated bellows<br />

mechanism of basidiospore release through<br />

an apical pore in puffballs and earth stars has<br />

evolved at least three times, and that the<br />

peridioles in the bird’s nest fungi (Cyathus,<br />

Crucibulum), in Sphaerobolus and in Pisolithus are<br />

analogous rather than homologous structures.<br />

Referring <strong>to</strong> these and other findings made by<br />

molecular phylogeneticists, Reijnders (2000) concluded<br />

that ‘if this key denotes real affinities,<br />

morphologists must be ashamed of their wrong<br />

conclusions’.<br />

Ingold (1971) regarded the gasteromycetes<br />

as a biological group which, having lost the<br />

active spore discharge mechanism of their<br />

hymenomycete ances<strong>to</strong>rs, have attempted a<br />

remarkable series of experiments in spore<br />

liberation. In order <strong>to</strong> explore this aspect of<br />

gasteromycete biology, we shall consider these<br />

fungi <strong>to</strong>gether in the present chapter, but<br />

drawing on the taxonomic framework as set<br />

out in Chapter 19.


580 HOMOBASIDIOMYCETES: GASTEROMYCETES<br />

Table 20.1. Taxonomic affinities and spore release mechanisms of selected gasteromycetes.<br />

Hymenomycete<br />

grouping<br />

Gasteromycete<br />

genus<br />

Gasterocarp<br />

type<br />

Propagule Dispersal Ecology<br />

Euagarics clade Bovista,Calvatia Puffball<br />

(epigeous)<br />

Lycoperdon Puffball<br />

(epigeous)<br />

Crucibulum, Bird’s nest<br />

Cyathus (epigeous)<br />

Nia<br />

Gelatinous<br />

(epigeous)<br />

Bole<strong>to</strong>id clade Rhizopogon False truffle<br />

(hypogeous)<br />

Melanogaster False truffle<br />

(hypogeous)<br />

Pisolithus Earth ball<br />

(epigeous)<br />

Scleroderma Earth ball<br />

(epigeous)<br />

Astraeus Earth star<br />

(epigeous)<br />

Calos<strong>to</strong>ma Puffball<br />

(epigeous)<br />

Gomphoid<br />

phalloid clade<br />

Geastrum<br />

Anthurus,<br />

Clathrus,<br />

Phallus,<br />

Mutinus<br />

Sphaerobolus<br />

Earth star<br />

(epigeous)<br />

Stinkhorn, etc.<br />

(epigeous)<br />

Cannonball<br />

(epigeous)<br />

Basidiospore A Saprotrophic<br />

Basidiospore B Saprotrophic<br />

Peridiole C Saprotrophic<br />

Basidiospore D Saprotrophic<br />

Basidiospore E Ec<strong>to</strong>mycorrhizal<br />

Basidiospore E Ec<strong>to</strong>mycorrhizal<br />

Peridiole or A Ec<strong>to</strong>mycorrhizal<br />

basidiospore<br />

Basidiospore A Ec<strong>to</strong>mycorrhizal<br />

Basidiospore B Ec<strong>to</strong>mycorrhizal<br />

Basidiospore B Saprotrophic<br />

Basidiospore B Saprotrophic<br />

Basidiospore F Saprotrophic<br />

Peridiole G Saprotrophic<br />

Dispersal mechanisms of propagules are as follows:<br />

A.Disintegration of gasterocarp followed by release of propagules by wind or animal trampling.<br />

B.Puffing through a pore after a raindrop hits the endoperidium (‘bellows mechanism’).<br />

C. Splash cup dispersal following impact by a raindrop.<br />

D. Passive release in<strong>to</strong> water.<br />

E.Distribution by burrowing animals and/or passive release in<strong>to</strong> the soil following disintegration of the<br />

gasterocarp.<br />

F.Insectdispersal following olfac<strong>to</strong>ry and visual attraction.<br />

G. Active discharge of peridiole by tension-snap mechanism.


GASTEROMYCETES IN THE EUAGARICS CLADE<br />

581<br />

20.3 Gasteromycetes in the<br />

euagarics clade<br />

The euagarics clade contains some 10 000 fungi<br />

in 26 families (Hibbett & Thorn, 2001; Kirk et al.,<br />

2001). Hymenia may be produced on the gills,<br />

pores and ridges of mushrooms and on the<br />

surface of coral-shaped fruit bodies, or basidia<br />

may be enclosed in gasterocarps. Among the<br />

gasteromycetes found within the euagarics,<br />

the two most important families are the<br />

Lycoperdaceae comprising puffballs and related<br />

forms, and the Nidulariaceae (bird’s nest fungi).<br />

20.3.1 Lycoperdaceae: puffballs<br />

Puffballs such as Lycoperdon, Vascellum and<br />

Calvatia form a phylogenetically well-defined<br />

group which seems <strong>to</strong> be closely related <strong>to</strong> the<br />

genus Macrolepiota both on the basis of DNA<br />

sequence analyses (Krüger et al., 2001) and<br />

because of similarities in the on<strong>to</strong>geny and<br />

architecture of rhizomorphs (Agerer, 2002). The<br />

current family Lycoperdaceae contains 18 genera<br />

and 158 species of gasteromycetes with epigeous<br />

fruit bodies. The mature gasterocarp is thinwalled<br />

and either forms an apical pore<br />

(in Lycoperdon) or disintegrates from the apex<br />

downwards (e.g. in Calvatia, Vascellum, Bovista).<br />

Basidiospores are brown in colour and have<br />

warty or spiny walls, with the distal part of the<br />

basidial sterigma often remaining attached <strong>to</strong><br />

mature spores (Portman et al., 1997). Most species<br />

are saprotrophic on soil and humus.<br />

Lycoperdon<br />

About 50 species are known, producing fruit<br />

bodies which are pear-shaped or <strong>to</strong>p-shaped.<br />

Most species grow on the ground. Lycoperdon<br />

pyriforme (Fig. 20.3) is unusual in growing directly<br />

on old stumps, rotting wood and sawdust heaps.<br />

It is not closely related <strong>to</strong> other Lycoperdon spp.<br />

and is now called Morganella by some authors.<br />

Gasterocarps of Lycoperdon spp. commonly arise<br />

on mycelial cords. The individual cells of the<br />

mycelium usually contain paired nuclei, but<br />

clamp connections are absent (Dowding &<br />

Bulmer, 1964). A longitudinal section of a<br />

young gasterocarp of L. pyriforme (Figs. 20.3a,b)<br />

shows that it is surrounded by a two-layered<br />

peridium, but as the fruit body expands the<br />

pseudoparenchyma<strong>to</strong>us exoperidium may slough<br />

off or crack in<strong>to</strong> numerous scales or warts<br />

(Fig. 20.2a) whilst the <strong>to</strong>ugher endoperidium<br />

made up of both thick-walled and thin-walled<br />

hyphae remains unbroken, apart from a pore at<br />

the apex of the fruit body. The tissue within the<br />

peridium is differentiated in<strong>to</strong> a non-sporing<br />

region or sub-gleba at the base of the gasterocarp,<br />

which extends as a columella in<strong>to</strong> the<br />

sporulating region (gleba) in the upper part of<br />

the fruit body. The glebal tissue is sponge-like,<br />

containing numerous small cavities, and in the<br />

upper fertile part the cavities are lined by the<br />

hymenium. The tissue separating the hymenial<br />

chambers is made up of thick- and thin-walled<br />

hyphae. The thin-walled hyphae break down as<br />

the gasterocarp ripens, but the thick-walled<br />

hyphae persist <strong>to</strong> form the capillitium threads<br />

between which the spores are contained. The<br />

basidia lining the cavities of the gleba are<br />

rounded and bear one <strong>to</strong> four basidiospores<br />

symmetrically arranged on sterigmata of varying<br />

length (Fig. 20.3c). Young basidia are binucleate,<br />

and nuclear fusion and meiosis occur in the<br />

usual way. One nucleus migrates in<strong>to</strong> each spore,<br />

and if fewer than four spores are produced,<br />

the spare nuclei degenerate in the basidium<br />

(Dowding & Bulmer, 1964). The basidiospores are<br />

not violently projected from the sterigmata. As<br />

the glebal tissue breaks down and dries, the<br />

spores are left as a brown dusty mass inside the<br />

fruit body. An apical pore develops by controlled<br />

lysis of the endoperidium (Fig. 20.2b). The thin<br />

upper layer of the endoperidium is elastic and<br />

acts as a bellows, and when rain drops impinge<br />

on this layer, small clouds of spores are puffed<br />

out (Gregory, 1949). Little is known of the mating<br />

behaviour of Lycoperdon.<br />

Calvatia<br />

Gasterocarps about the size of a rugby football<br />

are produced by Calvatia (Langermannia) gigantea<br />

(Fig. 20.2c) growing on grassland and on<br />

disturbed ground. There is no definite pore; the<br />

peridium breaks away <strong>to</strong> expose a brown spore<br />

mass. Buller (1909) estimated that the output of


582 HOMOBASIDIOMYCETES: GASTEROMYCETES<br />

Fig 20.2 Gasteromycetes in the euagarics clade. (a,b) Lycoperdon perlatum.Young gasterocarps (a) are ornamented by warts formed<br />

from the exoperidium. In older gasterocarps (b), the warts have sloughed off and the endoperidium has formed the apical pore.<br />

(c) Calvatia gigantea.Young fruit bodies growing with Urtica dioica.The coin is 2.5 cm in diameter. (d) Cyathus stercoreus.Gasterocarps<br />

(about 5 mm diameter) opening up <strong>to</strong> reveal the peridioles.<br />

a specimen measuring 40 28 20 cm was<br />

7 10 12 spores (see Table 18.1), and much larger<br />

specimens have been recorded. The spores are<br />

spherical with scattered warts. A more<br />

commonly encountered species is C. excipuliformis,<br />

which fruits on humus (Plate 11a).<br />

When attempts are made <strong>to</strong> germinate<br />

the spores of this and other puffballs in<br />

the labora<strong>to</strong>ry, the percentage of germination is<br />

extremely low, often less than 0.1%. Germination<br />

takes several weeks and is stimulated by the<br />

growth of yeasts (Bulmer, 1964; Wilson & Beneke,<br />

1966).<br />

20.3.2 Nidulariaceae: bird’s nest fungi<br />

Here the gasterocarps are funnel-shaped, and<br />

the gleba is differentiated in<strong>to</strong> lens-shaped<br />

peridioles (glebal masses) which contain the<br />

basidiospores. Some 50 species in 4 genera are<br />

known, of which the most common examples<br />

are Cyathus and Crucibulum. Detailed and highly<br />

readable accounts of the biology of bird’s nest<br />

fungi have been given by Brodie (1975, 1984).<br />

Members of this family are saprotrophic and<br />

are capable of degrading lignin (Wicklow et al.,<br />

1984).<br />

Cyathus<br />

The fruit bodies of C. olla can be found in autumn<br />

growing amongst cereal stubble. Cyathus striatus,<br />

recognized by the furrowed inner wall of its<br />

cups, grows on old stumps and twigs whilst<br />

C. stercoreus grows on old dung patches. This last<br />

species can be made <strong>to</strong> fruit readily if mycelium


GASTEROMYCETES IN THE EUAGARICS CLADE<br />

583<br />

Fig 20.3 Lycoperdon pyriforme. (a) L.S. gasterocarp. (b) Portion of peridium and gleba. Note the pseudoparenchyma<strong>to</strong>us<br />

exoperidium and the fibrous endoperidium. (c) Portion of gleba showing basidia, thin-walled hyphae and capillitium threads.<br />

grown on a mixture of cow dung and straw<br />

is covered with a thin layer of casing soil and<br />

then left at room temperature for a few weeks<br />

(Fig. 20.2d; Webster & Weber, 1997). The fungus<br />

can also fruit on agar media (Lu, 1973). Light is<br />

essential for fruit body formation. Since the<br />

peridioles retain viability for many years if<br />

frozen, C. stercoreus provides a convenient example<br />

<strong>to</strong> study.<br />

The first sign of fruit body development is the<br />

appearance of brown mycelial cords at the soil<br />

surface, on which knots of hyphae differentiate.<br />

In young gasterocarps, the mouth of the funnel<br />

is closed over by a thin papery epiphragm<br />

(Figs. 20.2d, 20.4a) which ruptures as the fruit<br />

body expands. Within the funnel, the peridioles<br />

develop. They are lens-shaped, slate-blue in<br />

colour and attached <strong>to</strong> the peridium by a<br />

complex funiculus. In earlier stages of development<br />

in this and other species of Cyathus, the<br />

peridioles are separated by thin-walled hyphae<br />

which disappear at maturity (Walker, 1920). The<br />

peridiole wall consists of an outermost layer<br />

(tunica) made up of loosely interwoven hyphae,<br />

a dark cortex, and an inner layer of thick-walled<br />

but hyaline cells (Fig. 20.4g). The fertile centre of<br />

the peridiole is made up of thin-walled hyphae<br />

(‘nurse hyphae’) between which basidia develop.<br />

The basidia form 4 8 basidiospores and disappear<br />

soon afterwards, but the spores continue <strong>to</strong><br />

enlarge and become thick-walled (Fig. 20.4g).<br />

Most of the nurse hyphae also break down,<br />

possibly providing nutrients for the enlarging<br />

spores.


584 HOMOBASIDIOMYCETES: GASTEROMYCETES<br />

Fig 20.4 Cyathus stercoreus. (a) Section of immature gasterocarp showing peridioles. (b) Gasterocarp cut open and pinned back <strong>to</strong><br />

show the attachment of the peridioles. (c e) Details of structure of funiculus. (c) Condition of funiculus before stretching.<br />

(d) Stretched funiculus.Note the funicular cord within the purse. (e) Funicular cord extended after rupture of the purse.The base of<br />

the funicular cord is frayed out <strong>to</strong> form the hapteron. (f) Portion of funicular cord. Note the spirally coiled hyphae.The thickenings<br />

are modified clamp connections. (g) Detail of peridiole wall and contents (b, basidiospore; c, cortex; e, epiphragm; em,<br />

emplacement;f,funiculus;f.c.,funicularcord;h,hapteron;m.c.,mylialcords;m.p.,middlepiece;p,peridiole;pu,purse;s,sheath;<br />

t, tunica).<br />

The gasterocarps of Cyathus and Crucibulum<br />

have been aptly termed splash cups because the<br />

peridioles are splashed out by the action of rain<br />

drops <strong>to</strong> distances of over 1 m. The key <strong>to</strong><br />

understanding the mechanism of discharge lies<br />

in the structure of the funiculus. In Cyathus<br />

(Figs. 20.4d,e), the funiculus is made up of several<br />

characteristic structures (see Brodie, 1975).<br />

The sheath is a tubular network of hyphae<br />

attached <strong>to</strong> the inner surface of the gasterocarp.<br />

It terminates in the middle piece where the<br />

innermost hyphae of the sheath unite <strong>to</strong> form a<br />

short cord. The middle piece flares out at its <strong>to</strong>p<br />

where its hyphae are attached <strong>to</strong> a cylindrical sac,<br />

the purse, which is firmly attached <strong>to</strong> the<br />

peridiole at a small depression. Folded up<br />

within the purse is a long strand of spirally<br />

coiled hyphae, the funicular cord (Fig. 20.4f). The<br />

free end of the funicular cord is composed of a<br />

tangled mass of adhesive hyphae, the hapteron.<br />

Rain drops, which may be as much as 4 mm in<br />

diameter and have a terminal velocity of about<br />

8ms 1 , fall in<strong>to</strong> the cup. Drops of this size<br />

are most likely <strong>to</strong> drip from the woodland<br />

canopy (Savile & Hayhoe, 1978). The force creates<br />

a strong upward thrust which tears open the


GASTEROMYCETES IN THE BOLETOID CLADE<br />

585<br />

purse. The funicular cord, spirally coiled up<br />

within the purse, swells explosively, stretching <strong>to</strong><br />

a length of about 2 3mm in C. stercoreus, whilst<br />

in C. striatus it may be as long as 4 12 cm. As the<br />

peridiole is flicked away, the sticky hapteron<br />

at the base of the funicular cord helps <strong>to</strong><br />

attach the peridiole <strong>to</strong> surrounding vegetation,<br />

and the momentum of the peridiole may cause<br />

the funicular cord <strong>to</strong> wrap around objects.<br />

Peridioles of C. stercoreus are presumably eaten<br />

by herbivorous animals and it is known that the<br />

basidiospores on release from the peridiole are<br />

stimulated <strong>to</strong> germinate by incubation around<br />

37°C. Whether animals play a significant role<br />

in the dispersal of other bird’s nest fungi is<br />

uncertain. The funiculus of Crucibulum is different<br />

from that of Cyathus, with a longer middle<br />

piece, a very short purse, and a funicular cord<br />

which is composed of relatively few hyphae only<br />

slightly coiled.<br />

Both Cyathus and Crucibulum show tetrapolar<br />

heterothallism with relatively few alleles<br />

(generally not more than 15) at each locus, and<br />

this is the most usual condition within the<br />

Nidulariaceae (Burnett & Boulter, 1963; Lu, 1964).<br />

20.4 Gasteromycetes in the<br />

bole<strong>to</strong>id clade<br />

The bole<strong>to</strong>id clade, as defined by molecular<br />

phylogeny (Hibbett & Thorn, 2001), contains<br />

fungi with a wide range of fruit body types,<br />

including lamellate (e.g. Paxillus), bole<strong>to</strong>id<br />

(e.g. Boletus, Leccinum, Suillus, Xerocomus) and<br />

resupinate forms (e.g. Coniophora, Serpula). These<br />

have been described previously (Section 19.5).<br />

Gasteromycete fungi have arisen from bole<strong>to</strong>id<br />

ances<strong>to</strong>rs on several occasions, with the family<br />

Sclerodermataceae having an affinity with<br />

Gyroporus in a ‘bole<strong>to</strong>id’ branch, and the<br />

Rhizopogonaceae with Suillus (‘suilloid’ branch;<br />

Grubisha et al., 2001; Binder & Bresinsky, 2002).<br />

Both families contain mainly ec<strong>to</strong>mycorrhizal<br />

fungi (see Table 20.1).<br />

Features of their association with tree roots<br />

are typical of members of the bole<strong>to</strong>id clade in<br />

that there is a large amount of fungal biomass<br />

extending from the mantle in<strong>to</strong> the soil by<br />

means of mycelial cords or rhizomorphs which<br />

may be several metres long. This type of<br />

ec<strong>to</strong>mycorrhiza appears <strong>to</strong> be particularly effective<br />

in exploiting a large volume of soil for<br />

nutrients, and it is also credited with improving<br />

the water status and thus the performance of the<br />

tree host under conditions of drought (Smith &<br />

Read, 1997; Agerer, 2001). The ability <strong>to</strong> form an<br />

extensive rhizomorph system may explain why<br />

mycorrhizal gasteromycetes belonging <strong>to</strong> the<br />

bole<strong>to</strong>id clade are particularly prominent in dry<br />

habitats. It is probable that long-distance transport<br />

processes in rhizomorphs are facilitated by<br />

peristaltic movement through a system of tubular<br />

vacuoles (Fig. 1.9; Ashford & Allaway,<br />

2002). Certain ec<strong>to</strong>mycorrhizal fungi such as<br />

Rhizopogon, Scleroderma and Pisolithus can be<br />

grown in pure culture, and basidiospore inoculum<br />

from their relatively large fruit bodies is also<br />

easily collected and s<strong>to</strong>red. Hence, these species<br />

are suitable for labora<strong>to</strong>ry-based research as well<br />

as inoculation of trees prior <strong>to</strong> outplanting in<strong>to</strong><br />

forestry situations.<br />

In addition <strong>to</strong> the morphology of ec<strong>to</strong>mycorrhiza,<br />

there are several further features betraying<br />

an affinity of the Sclerodermataceae and<br />

Rhizopogonaceae with the bole<strong>to</strong>id clade. For<br />

instance, pulvinic acid-type pigments typical of<br />

Boletus, Suillus and Xerocomus (see Fig. 19.22) are<br />

also found in their gasteromycete relatives,<br />

either in a pure form (e.g. variegatic acid,<br />

xerocomic acid) or as derivatives (Gill &<br />

Watling, 1986; Gill & Steglich, 1987; Winner<br />

et al., 2004). Further, the mycoparasitic mould<br />

Apiocrea chrysosperma (anamorph Sepedonium<br />

chrysospermum), which frequently forms a<br />

golden yellow conidial crust on fruit bodies of<br />

Boletus, Suillus, Xerocomus and Paxillus (Plate 9h),<br />

also infects gasteromycetes such as Scleroderma<br />

and Rhizopogon (see Gill & Watling, 1986).<br />

Pathogens may be competent taxonomists!<br />

20.4.1 Sclerodermataceae: earth balls and<br />

relatives<br />

This family comprises some 50 species in<br />

7 genera (Kirk et al., 2001), including the earth<br />

ball Scleroderma, the stalked puffball Calos<strong>to</strong>ma,


586 HOMOBASIDIOMYCETES: GASTEROMYCETES<br />

the dye ball Pisolithus, and the barometer<br />

earth star Astraeus. Binder and Bresinsky (2002)<br />

have examined the phylogeny of the group<br />

and have recommended its partitioning in<strong>to</strong><br />

several families. All members seem <strong>to</strong> be ec<strong>to</strong>mycorrhizal<br />

with trees. Here we shall consider<br />

the two most important genera, Pisolithus and<br />

Scleroderma.<br />

Pisolithus<br />

The best-known species is P. tinc<strong>to</strong>rius (¼<br />

P. arhizus), which is so called because its<br />

immature gasterocarps, when injured, produce<br />

an intense black dye (see Plate 11b). Because of<br />

the variability of gasterocarp appearance, the<br />

taxonomy of Pisolithus has been problematic, and<br />

initially P. tinc<strong>to</strong>rius was thought <strong>to</strong> be of panglobal<br />

distribution, capable of associating with<br />

almost any ec<strong>to</strong>mycorrhiza-forming tree species<br />

(Marx, 1977). Pisolithus is now known <strong>to</strong> consist of<br />

more than 10 species (Cairney, 2002; Martin et al.,<br />

2002), with P. tinc<strong>to</strong>rius distributed throughout<br />

the Northern Hemisphere and associated mainly<br />

with Pinus and Quercus. The centre of evolution of<br />

the genus is probably Australia, and P. marmoratus<br />

associated with Eucalyptus is regarded as the<br />

Southern Hemisphere equivalent of P. tinc<strong>to</strong>rius.<br />

This and other species have been spread <strong>to</strong> South<br />

America, South East Asia and Africa, <strong>to</strong>gether<br />

with their host trees (Eucalyptus, Acacia) which are<br />

used in intensive forestry and in reforestation<br />

programmes (Dell et al., 2002; Martin et al.,<br />

2002). In addition <strong>to</strong> this anthropogenic dispersal,<br />

there is also evidence that Pisolithus can travel<br />

long distances as air-borne basidiospores, e.g.<br />

from Australia <strong>to</strong> New Zealand (Moyersoen et al.,<br />

2003).<br />

Pisolithus spp. may be displaced by other<br />

ec<strong>to</strong>mycorrhizal fungi in cool, wet situations<br />

(McAfee & Fortin, 1986) but are prominent in<br />

extreme environments, e.g. dry habitats with<br />

sandy soil, or areas polluted with heavy metals<br />

(Walker et al., 1989; Smith & Read, 1997). In such<br />

situations, the growth of mycorrhizal trees can<br />

be increased several-fold relative <strong>to</strong> uninoculated<br />

controls. Benefits of Pisolithus infections <strong>to</strong> the<br />

host tree include enhanced provision of nutrients<br />

and water, de<strong>to</strong>xification of heavy<br />

metals, and protection against soil-borne plant<br />

pathogens. The considerable promise of Pisolithus<br />

is reflected by an immense body of literature<br />

which has been summarized admirably by<br />

Cairney and Chambers (1997) and Chambers<br />

and Cairney (1999).<br />

Pisolithus has a tetrapolar mating system<br />

(Kope & Fortin, 1990), and although monokaryons<br />

can infect tree roots, a full-scale ec<strong>to</strong>mycorrhizal<br />

association requires a dikaryotic<br />

mycelium. The establishment of a mycorrhiza<br />

proceeds in several steps. Chemotropic growth of<br />

Pisolithus hyphae <strong>to</strong>wards host root tips is<br />

followed by the secretion of glycoprotein fibrils<br />

by the fungus during initial contact (Lei et al.,<br />

1990). Dead or moribund cells in the root cap<br />

region are infected first; the mantle is then<br />

established within 48 h, and a Hartig net formed<br />

subsequently (Horan et al., 1988; Lei et al., 1990).<br />

Only root material grown after initial contact is<br />

colonized. Rhizomorphs radiate outwards for<br />

several metres, and these may partly account<br />

for the success of the Pisolithus mycorrhiza,<br />

especially in dry habitats. Another fac<strong>to</strong>r may<br />

be the formation of sclerotia which enable the<br />

fungus <strong>to</strong> survive adverse conditions in the soil<br />

(Grenville et al., 1985). Eventually, fruit body<br />

initials are formed in the soil, with the maturing<br />

gasterocarps pushing through the surface.<br />

Basidiospores are produced inside numerous<br />

peridioles which disintegrate <strong>to</strong> release their<br />

spores passively (Plate 11b). The formation and<br />

maturation of peridioles proceeds from the tip<br />

<strong>to</strong> the base of the gasterocarp which gradually<br />

breaks up in the process. Gasterocarps can be<br />

sizeable, up <strong>to</strong> 20 cm tall.<br />

Studies on Pisolithus mycelia in Australian<br />

eucalypt forests have revealed that genetically<br />

distinguishable individuals (genets) may be<br />

variable in size, ranging from less than 2 m 2 <strong>to</strong><br />

50 m 2 or more. Since these are interspersed,<br />

the smaller genets are interpreted as the result<br />

of recent re-colonization events from winddispersed<br />

basidiospore inoculum (Anderson<br />

et al., 1998, 2001).<br />

Scleroderma<br />

There are about 25 species of Scleroderma (Sims<br />

et al., 1995), three common temperate examples<br />

being S. bovista, S. citrinum (Fig. 20.5) and


GASTEROMYCETES IN THE BOLETOID CLADE<br />

587<br />

Fig 20.5 Scleroderma citrinum. (a) Maturing gasterocarps, about 3 6 cm in diameter.One has been cut open <strong>to</strong> reveal the gleba<br />

containing purplish-black basidiospores. (b) Old gasterocarps.The peridium has cracked open, permitting passive dispersal of the<br />

black basidiospore mass.<br />

In mature gasterocarps, the peridium is<br />

apparently a single, fairly thick layer. Although<br />

the glebal mass may be traversed by a system<br />

of sterile veins, there is no columella and<br />

no capillitium. The basidiospores are sessile<br />

(Fig. 20.6). When the gasterocarp is ripe, it<br />

cracks open irregularly and the dry spores<br />

escape (Fig. 20.5b). There is no well-developed<br />

bellows mechanism as in Lycoperdon (p. 579) or<br />

Geastrum (p. 588).<br />

Fig 20.6 Scleroderma verrucosum. Basidia and basidiospores.<br />

Note that the spores are almost sessile.<br />

S. verrucosum (Pegler et al., 1995). Earth balls are<br />

found in the autumn in acid woodlands and<br />

heaths under such trees as Pinus, Betula, Quercus<br />

and Fagus with which they form ec<strong>to</strong>mycorrhizal<br />

associations. Mycorrhizal infection is easily<br />

reproduced under labora<strong>to</strong>ry conditions, using<br />

aqueous spore suspensions (Parladé et al.,<br />

1996). General aspects of mycorrhiza involving<br />

Scleroderma have been reviewed by Jeffries (1999).<br />

20.4.2 Rhizopogonaceae: beard truffles<br />

This family comprises some 150 species in<br />

4 genera. By far the most important genus is<br />

Rhizopogon, which is mycorrhizal mostly<br />

with coniferous trees. It originates from the<br />

Northern Hemisphere, with an unusually high<br />

diversity of species encountered in the Pacific<br />

Northwest where several host trees are native<br />

(Martín, 1996; Molina et al., 1999). Rhizopogon is<br />

related <strong>to</strong> Suillus in the bole<strong>to</strong>id clade (Grubisha<br />

et al., 2001).<br />

Rhizopogon<br />

Members of this genus form gasterocarps which<br />

resemble those of Scleroderma but are hypogeous,<br />

arising from mycelial cords (Plate 11c).<br />

The gasterocarps may be eaten by burrowing


588 HOMOBASIDIOMYCETES: GASTEROMYCETES<br />

mammals, and basidiospores pass through their<br />

digestive tracts unharmed (Colgan & Claridge,<br />

2002). Rhizopogon luteolus and R. roseolus are<br />

common and cosmopolitan under Pinus spp.,<br />

whereas R. vinicolor and others are specific<br />

associates of Douglas fir (Pseudotsuga menziesii).<br />

Reviews of the genus have been written<br />

by Molina and Trappe (1994) and Molina<br />

et al. (1999).<br />

Rhizopogon spp. provide similar benefits <strong>to</strong><br />

their hosts as Pisolithus and are the subject of<br />

research and development activities in forestry.<br />

Inoculation is readily achieved by dusting seeds<br />

with basidiospores or immersing the roots of<br />

seedlings in spore suspensions (Parladé et al.,<br />

1996). Although Rhizopogon spp. can be grown in<br />

pure culture, inoculation of host tree seedlings<br />

with mycelium is not generally efficient with<br />

this and related fungi, unless the hyphae are<br />

protected from mechanical damage, for example,<br />

by being grown inside porous or gel-like beads<br />

(see Smith & Read, 1997).<br />

20.5 Gasteromycetes in the<br />

gomphoid phalloid clade<br />

The gomphoid phalloid clade contains some<br />

350 species of morphologically diverse fungi<br />

(Hibbett & Thorn, 2001). Most of the species<br />

with active basidiospore discharge form coral- or<br />

club-shaped basidiocarps, e.g. Ramaria (see Plate<br />

10h), Clavariadelphus and Gomphus. Several groups<br />

of well-known gasteromycetes also belong here<br />

(Hibbett et al., 1997b; Humpert et al., 2001), and<br />

these show the most spectacular spore dispersal<br />

mechanisms of all gasteromycetes, including<br />

the bellows mechanism, insect dispersal, and<br />

active discharge of peridioles. Details of the<br />

phylogeny or evolutionary his<strong>to</strong>ry of these<br />

gasteromycetes still appear <strong>to</strong> be unknown,<br />

although some groups such as earth stars may<br />

be ancient (Krassilov & Makulbekov, 2003).<br />

Members of the gomphoid phalloid clade are<br />

mostly saprotrophic on wood, other plant debris<br />

and humus, extending <strong>to</strong> the soil by means of<br />

mycelial cords.<br />

20.5.1 Geastraceae: earth stars<br />

There are some 50 species of earth stars.<br />

Descriptions of the common species may be<br />

found in the keys by Ellis and Ellis (1990) and<br />

Pegler et al. (1995). One of the most frequent and<br />

widespread species in temperate and subtropical<br />

forests is Geastrum triplex (Fig. 20.7a) which grows<br />

in the leaf litter of beech, sycamore and pine. The<br />

young fruit body is onion-shaped and develops at<br />

or just below the soil surface. The exoperidium is<br />

complex, consisting of a brown outer layer made<br />

of narrow hyphae mostly running longitudinally,<br />

and a paler pseudoparenchyma<strong>to</strong>us inner<br />

layer. As the fruit body ripens, the whole of the<br />

exoperidium splits open from the tip in a stellate<br />

fashion and, due <strong>to</strong> swelling of the pseudoparenchyma<br />

cells of the exoperidium, the triangular<br />

flaps curve outwards and make contact with the<br />

soil, lifting the inner part of the fruit body in<strong>to</strong><br />

the air (Fig. 20.7a). The thin, papery endoperidium<br />

opens by an apical pore. Spores are puffed<br />

out by the bellows mechanism when rain drops<br />

strike the endoperidium (Ingold, 1971). The<br />

gleba contains a columella (sometimes termed<br />

a pseudocolumella) and capillitium, much as in<br />

the puffball Lycoperdon (Fig. 20.3) with which<br />

Geastrum is not related. Basidial development can<br />

only be observed in young unexpanded gasterocarps.<br />

The basidia are pear-shaped, with 4 6<br />

(sometimes up <strong>to</strong> 8) spores borne on a knob-like<br />

extension of the pointed end.<br />

20.5.2 Sphaerobolus<br />

Sphaerobolus is unique among gasteromycetes in<br />

that it has developed an active discharge<br />

mechanism of peridioles (glebal masses), thereby<br />

reversing the loss of active basidiospore liberation.<br />

The precise taxonomic position of<br />

Sphaerobolus is still unclear at present; formerly<br />

grouped <strong>to</strong>gether with the bird’s nest fungi<br />

(p. 578), it is now known <strong>to</strong> belong <strong>to</strong> the<br />

gomphoid phalloid clade. Kirk et al. (2001) have<br />

included it in the Geastraceae.<br />

Sphaerobolus stellatus forms globose orange<br />

gasterocarps about 2 mm in diameter. They are<br />

attached <strong>to</strong> rotten wood, rotting herbaceous<br />

stems, sacking and weathered dung of herbivores<br />

such as cow and sheep. Ripe fruit bodies open


GASTEROMYCETES IN THE GOMPHOID PHALLOID CLADE<br />

589<br />

Fig 20.7 Gasteromycetes belonging <strong>to</strong> the gomphoid phalloid clade. (a) The earth star Geastrum triplex.(b)Sphaerobolus stellatus.<br />

Stages of maturation can be seen from left (immature gasterocarp) through the centre (two opened gasterocarps exhibiting glebal<br />

masses) <strong>to</strong> right (discharged gasterocarp with everted inner cup). (c) The veiled stinkhorn, Phallus indusiatus.This beautiful species is<br />

called ‘queen of mushrooms’ (kinoko no joou) in Japan. (d) The dog’s stinkhorn, Mutinus caninus. (b) reproduced from Webster and<br />

Weber (1999), with permission from Elsevier.Original print of (c) kindly provided by N.Tuno.<br />

<strong>to</strong> form a star-like arrangement of two cups<br />

fitting inside each other, attached only by the<br />

triangular tips of their teeth (Fig. 20.7b). Within<br />

the inner cup is a single brown peridiole or<br />

glebal mass about 1 mm in diameter. By sudden<br />

eversion of the inner cup, the peridiole is<br />

projected for a considerable distance. Buller<br />

(1933) has given a detailed account of peridiole<br />

discharge. He showed that the peridiole could be<br />

projected vertically for more than 2 m, and<br />

horizontally for over 4 m, with the record<br />

currently standing at 5.7 m (Ingold, 1971, 1972).<br />

The fungus can be cultivated if a peridiole is<br />

placed in a plate of oatmeal agar, and gasterocarps<br />

are produced after a few weeks’ incubation<br />

in daylight on this medium or on chopped<br />

straw saturated with a nutrient solution (Flegler,<br />

1984; Webster & Weber, 1999).<br />

A section through an almost mature, but<br />

unopened, gasterocarp is shown in Fig. 20.8a. The<br />

peridiole is surrounded by a peridium in which<br />

six layers can be distinguished. Three of these<br />

layers form the structure of the outer cup. The<br />

three layers making up the inner cup consist of<br />

an outer layer of tangentially arranged interwoven<br />

hyphae, a central layer of radially elongated<br />

cells forming a kind of palisade, and a thin<br />

innermost layer of pseudoparenchyma whose<br />

cells undergo deliquescence before glebal<br />

discharge <strong>to</strong> form a liquid which bathes the<br />

gleba and lies in the bot<strong>to</strong>m on the inner cup.<br />

Before the gasterocarp opens up, the cells of the<br />

palisade layer are rich in glycogen, but this is<br />

converted <strong>to</strong> glucose during ripening (Engel &<br />

Schneider, 1963). Intracellular accumulation of<br />

glucose causes the osmotic concentration of<br />

the cells <strong>to</strong> rise so that they absorb water and<br />

become more turgid. The swelling of the palisade<br />

layer is restrained by the tangentially arranged<br />

hyphae, and this sets up strains within the<br />

tissues of the inner cup which are only released<br />

by its turning inside out.<br />

Light is necessary for development, and the<br />

opening of the fruit body is pho<strong>to</strong>tropic,


590 HOMOBASIDIOMYCETES: GASTEROMYCETES<br />

Fig 20.8 Sphaerobolus stellatus. (a) V.S. of nearly ripe gasterocarp showing the central glebal mass (peridiole) surrounded by a<br />

six-layered peridium. (b) Details of the peridial layers:1, outermost layer composed of interwoven hyphae; 2, layer in which the<br />

hyphae are separated by extensive mucilage; 3, pseudoparenchyma<strong>to</strong>us layer; 4, fibrous layer; 5, palisade layer; 6, layer of lubricating<br />

cells; gl, outer layers of glebal mass. (c) Enlarged portion of layers 4 6 and portion of the glebal mass: c, cystidia; g, gemmae;<br />

b, basidiospores. (d) Clusters of basidia from unripe gasterocarps.There are usually 4 6 basidiospores. (e) Gemmae.<br />

(f) Basidiospores. (c, e) and (f) <strong>to</strong> same scale.<br />

ensuring that the glebal mass is projected<br />

<strong>to</strong>wards the light (Alasoadura, 1963). Peridiole<br />

discharge follows a diurnal rhythm, with<br />

release occurring during the light phase. In<br />

continuous light rhythmic discharge ceases, but<br />

in continuous darkness a culture previously<br />

exposed <strong>to</strong> alternating periods of 12 h of light<br />

and darkness continues <strong>to</strong> discharge peridioles<br />

rhythmically at times corresponding <strong>to</strong> the<br />

previous light periods, indicating an endogenous<br />

circadian rhythm.<br />

The spherical peridiole (glebal mass) is<br />

surrounded by a dark brown sticky coat derived<br />

from the breakdown of the cells of the innermost<br />

peridial layer. Immediately within the brown<br />

outer coat of the glebal mass are layers of rounded<br />

cells sometimes termed cystidia (Fig. 20.8c).<br />

Apparently these cells are incapable of germination<br />

and their function is not known. The rest of<br />

the glebal mass consists of oval thick-walled<br />

haploid basidiospores and thinner-walled dikaryotic<br />

gemmae. About 4 8 basidiospores develop<br />

on the basidia some 2 days before discharge (Fig.<br />

20.8d), but the basidia disappear as the glebal<br />

mass ripens. Gemmae arise either terminally or in<br />

an intercalary position on hyphae within the<br />

glebal mass. Oil-rich cells are also present. The<br />

sticky peridiole adheres readily <strong>to</strong> objects on<br />

which it is impacted, and after drying it is very<br />

difficult <strong>to</strong> dislodge even by a jet of water.<br />

Peridioles are viable for several years. Projectiles<br />

adhering <strong>to</strong> herbage may be eaten by animals, and<br />

this may explain the presence of fruit bodies on<br />

dung.<br />

On germination the peridioles give rise <strong>to</strong><br />

clamped hyphae which usually arise directly from<br />

the gemmae and not from the basidiospores. Most<br />

basidiospores, if they germinate, give rise <strong>to</strong><br />

mycelia with simple septa. Pairings of monosporous<br />

mycelia have indicated that the fungus is<br />

usually heterothallic, although details of the<br />

mating system still appear unclear at present.<br />

20.5.3 Phallaceae: stinkhorns<br />

An original solution <strong>to</strong> the problem posed by the<br />

loss of active basidiospore discharge has been<br />

developed also by members of the Phallaceae<br />

which attract insects, especially cadaverfeeding<br />

flies such as bluebottles, <strong>to</strong> visit their<br />

gasterocarps. Attraction may be by the emission


GASTEROMYCETES IN THE GOMPHOID PHALLOID CLADE<br />

591<br />

of a cadaverous smell or by colour, with gasterocarps<br />

of species like Clathrus ruber (Plate 11e) and<br />

C. archeri (Plate 11f) appearing dark red due <strong>to</strong> the<br />

accumulation of carotenoids, chiefly lycopene<br />

(Fiasson & Petersen, 1973; Gill & Steglich, 1987).<br />

There may be a synergism of attractions because<br />

species with brightly coloured gasterocarps tend<br />

<strong>to</strong> emit less evil smells than dull-coloured ones.<br />

Common temperate examples of Phallaceae<br />

are the graphically named stinkhorn Phallus<br />

impudicus (Lat. impudicus ¼ shameless; Plate 11d)<br />

and the dog’s stinkhorn Mutinus caninus<br />

(Fig. 20.7d).<br />

Phallus<br />

In late summer and autumn, stinkhorns can<br />

be detected readily by their smell. They can be a<br />

common or even dominant component of the<br />

population of basidiocarps on the forest floor, as<br />

shown by Shorrocks and Charlesworth (1982) who<br />

estimated some 50 000 70 000 gasterocarps km 2<br />

per season in a woodland. In such situations,<br />

stinkhorns and their primordia, the ‘eggs’,<br />

provide a major breeding ground for mycophagous<br />

flies. Eggs of P. impudicus are about 5 cm in<br />

diameter and develop from an extensive system of<br />

white mycelial cords which can be traced underground<br />

<strong>to</strong> a buried tree stump (see Fig. 1.12b;<br />

Grainger, 1962). A longitudinal section of an egg<br />

(Fig. 20.9a) shows a thin papery outer and inner<br />

peridium and a wider mass of jelly making up the<br />

middle peridium. The central part of the gasterocarp<br />

is differentiated in<strong>to</strong> a cylindrical hollow<br />

stipe and a folded honeycomb-like receptacle<br />

which bears the fertile part of the gleba. Within<br />

the young gleba are cavities lined by basidia<br />

bearing up <strong>to</strong> nine spores (Fig. 20.9b), but as the<br />

glebal mass ripens the basidia disintegrate.<br />

Gasterocarps expand very rapidly: within a few<br />

hours the stipe may elongate from about 5 cm <strong>to</strong> a<br />

length of 15 cm or more, leaving behind the<br />

peridial remains as a volva at its base.<br />

A demonstration of gasterocarp erection in the<br />

labora<strong>to</strong>ry by incubation of freshly collected ripe<br />

eggs in a moist chamber rarely fails <strong>to</strong> impress.<br />

The sudden expansion is probably at the expense<br />

of water s<strong>to</strong>red within the jelly of the middle<br />

peridium. The mean weight of expanded stipes is<br />

more than twice that of unexpanded ones (Ingold,<br />

1959). Expansion of the stipe of P. impudicus is<br />

accompanied by breakdown of glycogen and its<br />

conversion <strong>to</strong> sugar (Buller, 1933). A similar<br />

conversion has been reported in P. indusiatus<br />

(Fig. 20.7c) in which cells of the unexpanded<br />

stipe are folded but expand <strong>to</strong> almost 12 times<br />

their original volume during stipe elongation<br />

(Kinugawa, 1965).<br />

At about the same time as the stipe of<br />

P. impudicus is elongating, the fertile glebal mass<br />

begins <strong>to</strong> release strong-smelling volatile<br />

substances which are attractive <strong>to</strong> flies, especially<br />

bluebottles (Plate 11d). Depending on the analytical<br />

methods used, the smell has been attributed<br />

<strong>to</strong> a range of substances including methylmercaptan<br />

and hydrogen sulphide (List & Freund,<br />

1968) or dimethyl disulphide (see Fig. 15.4) and<br />

dimethyl trisulphide (Borg-Karlson et al., 1994).<br />

Once a fly has located a gasterocarp, it is<br />

presented with the dark green glebal<br />

mass of basidiospores embedded in a liquid<br />

which contains sugars and also sweet-smelling<br />

substances such as phenylacetaldehyde and<br />

phenylethanol. Flies feed on the spore mass<br />

which is removed within a few hours, leaving<br />

behind the pale receptacle. The ingested basidiospores<br />

are defaecated, apparently unharmed,<br />

on<strong>to</strong> surrounding vegetation and the soil, often<br />

within a short time of ingestion. Tuno (1998)<br />

found that the gut of fruitflies (Drosophila spp.)<br />

feeding on P. indusiatus and P. duplicatus contained<br />

up <strong>to</strong> 240 000 basidiospores, and that of the larger<br />

muscid flies up <strong>to</strong> 1.7 million. Basidiospore<br />

germination was unaffected by passage through<br />

the gut. However, it is unknown how the<br />

mycelium from germinating basidiospores<br />

succeeds in reaching fresh tree stumps, and<br />

clonal spread by mycelial cords may be an<br />

important additional mode of reproduction.<br />

There have been few studies of the nutrition<br />

and physiology of Phallus, but P. ravenelii has<br />

been shown <strong>to</strong> make good vegetative growth on<br />

a wide range of carbohydrates, and <strong>to</strong> require<br />

thiamine (Howard & Bigelow, 1969). The veiled<br />

stinkhorns P. indusiatus (Fig. 20.7c) and P. duplicatus<br />

are grown in China as a culinary<br />

speciality. Expanded fruit bodies are produced<br />

commercially from inoculated wood both in<br />

woodlands and indoors, and are marketed in a


592 HOMOBASIDIOMYCETES: GASTEROMYCETES<br />

Fig 20.9 Phallus impudicus. (a) L.S.‘egg’ showing the<br />

unexpanded stipe. (b) Basidia.<br />

dried form; sensibly, the volva and glebal mass are<br />

removed prior <strong>to</strong> marketing (Chang & Miles,<br />

2004). Apparently, the egg stage is also eaten<br />

after it has been boiled. These species were<br />

formerly thought <strong>to</strong> belong <strong>to</strong> a separate genus,<br />

Dictyophora, on the basis of the presence of a veil.<br />

However, this character is no longer considered<br />

useful because even the common stinkhorn may<br />

occasionally produce a short veil and is then<br />

called P. impudicus var. <strong>to</strong>gatus (see Kibby &<br />

Bingham, 2004).<br />

Other genera<br />

The general form of Mutinus, the dog’s stinkhorn,<br />

is similar <strong>to</strong> Phallus, but the gasterocarps are<br />

smaller (Fig. 20.7d). The upper part of the stipe is<br />

orange in colour, and the smell is less overpowering.<br />

The receptacle bearing the glebal mass<br />

is not reticulate.<br />

Clathrus ruber (Plate 11e) forms conspicuous,<br />

red cage-like gasterocarps in warm, dry<br />

habitats in the Mediterranean, occasionally<br />

extending <strong>to</strong> temperate climates. Other Clathrus<br />

spp. are tropical. Clathrus archeri, the squid<br />

fungus, was probably introduced <strong>to</strong> Europe<br />

from Australia some 100 years ago<br />

(Ramsbot<strong>to</strong>m, 1953) and is now established<br />

in many localities. An even more bizarre<br />

Australian species is Aseroe rubra, the starfish<br />

fungus (Plate 11f).


21<br />

Heterobasidiomycetes<br />

21.1 <strong>Introduction</strong><br />

The class Heterobasidiomycetes is approximately<br />

synonymous with the terms ‘Phragmobasidiomycetes’<br />

or ‘jelly fungi’ and contains fungi with<br />

the following characteristics.<br />

1. The dolipore septum is complex, i.e. it is<br />

surrounded by a parenthesome. Parenthesomes<br />

are also found in the Homobasidiomycetes (Chapters<br />

19 and 20), but not in the Urediniomycetes<br />

(Chapter 22) and Ustilaginomycetes (Chapter 23).<br />

2. The basidia of Heterobasidiomycetes may<br />

be strongly lobed and often divided by transverse,<br />

oblique or longitudinal septa. Such basidia are<br />

loosely termed heterobasidia, especially if they<br />

arise directly from hyphae instead of teliospores<br />

as in most Urediniomycetes and Ustilaginomycetes.<br />

If the basidia are septate, they are also<br />

called phragmobasidia. The sterigma of the<br />

heterobasidium is unusually prominent and is<br />

often termed epibasidium (Martin, 1945). In<br />

contrast, the basidia of Homobasidiomycetes<br />

are club-shaped and always single-celled.<br />

3. The fruit bodies of most Heterobasidiomycetes<br />

are simpler in architecture than those of<br />

Homobasidiomycetes, and the hymenium is not<br />

normally protected by a roof- or shelf-like architecture.<br />

In compensation, these simple fruit<br />

bodies are generally able <strong>to</strong> survive drying and<br />

rehydration, with fresh crops of basidiospores<br />

produced after each rehydration event. Fully<br />

hydrated basidiocarps are typically greatly<br />

swollen and gelatinous, hence the term ‘jelly<br />

fungi’ for the Heterobasidiomycetes.<br />

4. The basidiospores of most species are<br />

capable of producing secondary spores which<br />

may be ballis<strong>to</strong>conidia, passively released conidia<br />

or yeast cells.<br />

Species included in this class show considerable<br />

morphological diversity, and taxonomic<br />

concepts have been in a state of flux. The first<br />

workers <strong>to</strong> emphasize the importance of basidial<br />

morphology were Pa<strong>to</strong>uillard (1887) and<br />

Brefeld (1888). Most orders currently included<br />

were placed here by Martin (1945) and Bandoni<br />

(1984), and these are listed in Table 21.1.<br />

The inclusion of these groups is supported<br />

by DNA-based phylogenetic studies (Weiss &<br />

Oberwinkler, 2001). The life cycles of Heterobasidiomycetes,<br />

as far as they are known, show<br />

an alternation of monokaryotic and dikaryotic<br />

stages. The two broad subclasses, Heterobasidiomycetidae<br />

and Tremellomycetidae, can be distinguished<br />

by their monokaryotic phase being<br />

mycelial or yeast-like, respectively. All heterobasidiomycete<br />

yeasts discussed in Chapter 24<br />

seem <strong>to</strong> belong <strong>to</strong> the Tremellomycetidae<br />

(Wells & Bandoni, 2001). Thorough and authoritative<br />

circumscriptions of the Heterobasidiomycetes<br />

have been written by Wells (1994) and Wells<br />

and Bandoni (2001).<br />

Ecologically, Heterobasidiomycetes are associated<br />

with wood and other decaying plant<br />

matter, either as saprotrophs or as mycoparasites<br />

of saprotrophic fungi. Some species, especially<br />

in the Cera<strong>to</strong>basidiales (Section. 21.2), play dual<br />

roles as necrotrophic pathogens of various plants<br />

and mycorrhizal associates of orchids. Heterobasidiomycetes<br />

are mainly terrestrial.


594 HETEROBASIDIOMYCETES<br />

Table 21.1. Orders and selected genera currently<br />

included in the Heterobasidiomycetes. After Wells<br />

and Bandoni (2001).<br />

Subclass Heterobasidiomycetidae<br />

1. Cera<strong>to</strong>basidiales (see below)<br />

Cera<strong>to</strong>basidium (anam. Rhizoc<strong>to</strong>nia ¼ Cera<strong>to</strong>rhiza).<br />

Thanatephorus (anam. Rhizoc<strong>to</strong>nia ¼ Moniliopsis).<br />

2. Tulasnellales (see Fig. 21.1b)<br />

Tulasnella (anam. Rhizoc<strong>to</strong>nia)<br />

3. Dacrymycetales (p. 598)<br />

Calocera<br />

Dacrymyces<br />

Ditiola<br />

4. Auriculariales (p. 601)<br />

Auricularia<br />

Exidia<br />

Pseudohydnum<br />

Sebacina<br />

SubclassTremellomycetidae<br />

1. Tremellales (p. 60 4)<br />

Cryp<strong>to</strong>coccus and Bullera (yeast forms;Table 24.1)<br />

Filobasidiella neoformans (see p.660)<br />

Tremella<br />

2. Christianseniales<br />

3. Filobasidiales (yeast forms; seeTable 24.1)<br />

4. Cys<strong>to</strong>filobasidiales (seeTable 24.1)<br />

Itersonilia (see p.493)<br />

Phaffia, Xanthophyllomyces (see p.665)<br />

5. Trichosporonales (yeastforms; seeTable 24.1)<br />

21.2 Cera<strong>to</strong>basidiales<br />

The most important members of this family<br />

belong <strong>to</strong> the anamorph genus Rhizoc<strong>to</strong>nia which<br />

we shall consider in detail. Sneh et al. (1996)<br />

and Roberts (1999) have compiled important<br />

reference works on this group. The secondary<br />

hyphae of Rhizoc<strong>to</strong>nia have conspicuous dolipore<br />

septa which are visible even with the light<br />

microscope (Tu et al., 1977). Electron microscopy<br />

studies have revealed the parenthesomes <strong>to</strong> be<br />

perforated by several large pores (Müller et al.,<br />

1998b). Clamp connections are not found in<br />

Rhizoc<strong>to</strong>nia but may be present in other members<br />

of the order. The hyphae of Rhizoc<strong>to</strong>nia are highly<br />

characteristic. Branches typically arise at a right<br />

angle <strong>to</strong> the leading hypha, with the branch<br />

point slightly constricted and a septum located a<br />

little way in<strong>to</strong> the branch (Fig. 21.2a). Depending<br />

on the species, the compartments of vegetative<br />

hyphae are binucleate or multinucleate; uninucleate<br />

hyphae are uncommon. Teleomorphic<br />

states are rare and conidia are not produced,<br />

but sclerotia of the loose type (see Fig. 1.16a) are<br />

frequently seen. The form-genus Rhizoc<strong>to</strong>nia has<br />

now been broken up in<strong>to</strong> several taxa which<br />

correlate with different teleomorphs (Moore,<br />

1987; Andersen & Stalpers, 1994). For instance,<br />

Moniliopsis has multinucleate hyphae and is<br />

referred <strong>to</strong> the teleomorph genus Thanatephorus<br />

whereas the hyphae of Cera<strong>to</strong>rhiza (teleomorph<br />

Cera<strong>to</strong>basidium) are binucleate (Tu et al., 1977;<br />

Vilgalys & Cubeta, 1994). <strong>Fungi</strong> resembling<br />

Rhizoc<strong>to</strong>nia transgress the boundaries of<br />

orders, with some teleomorphs referable <strong>to</strong> the<br />

Tulasnellales. Indeed, a few Rhizoc<strong>to</strong>nia-like fungi<br />

have even been assigned <strong>to</strong> the Ascomycota.<br />

Whereas hyphae of Rhizoc<strong>to</strong>nia are readily<br />

recognized as such, they offer few microscopic<br />

features for species identification, and a system<br />

based on anas<strong>to</strong>mosis groups, i.e. the ability of<br />

a given isolate <strong>to</strong> undergo plasmogamy with<br />

hyphae of defined tester strains, has been<br />

developed (Sneh et al., 1991). Such pairings may<br />

yield three different responses, with intermediate<br />

reactions also possible. (1) In genetically<br />

identical or closely related strains, anas<strong>to</strong>mosis<br />

leads <strong>to</strong> perfect fusion. (2) In less closely related<br />

members of the same anas<strong>to</strong>mosis group, anas<strong>to</strong>mosis<br />

is followed by death of the fusion cell due<br />

<strong>to</strong> vegetative incompatibility. (3) No anas<strong>to</strong>mosis<br />

occurs between members of different anas<strong>to</strong>mosis<br />

groups. The best-studied taxon, R. solani,<br />

contains about a dozen anas<strong>to</strong>mosis groups,<br />

some of which have been further divided in<strong>to</strong><br />

subgroups according <strong>to</strong> biochemical or other<br />

criteria. Although the individual anas<strong>to</strong>mosis<br />

groups within R. solani and other taxa (e.g.<br />

Cera<strong>to</strong>rhiza) correlate with phylogenetic clusters<br />

obtained by DNA-based phylogeny (Kuninaga<br />

et al., 1997; Gonzalez et al., 2001) and also <strong>to</strong><br />

a certain extent with the range of plant hosts


CERATOBASIDIALES<br />

595<br />

affected (Sneh et al., 1991), the question of the<br />

boundaries of biological species remains unanswerable<br />

at present.<br />

The basidiocarps of Cera<strong>to</strong>basidiales are thin,<br />

gelatinous and often resupinate. Their formation<br />

can sometimes be induced by covering an agar<br />

culture with soil (Warcup & Talbot, 1966). Hyphal<br />

tips due <strong>to</strong> develop in<strong>to</strong> a probasidium are<br />

generally binucleate, and karyogamy is followed<br />

swiftly by meiosis. During the later stages of<br />

meiosis, four prominent sterigmata referred <strong>to</strong><br />

as epibasidia are formed (Wells & Bandoni, 2001),<br />

but septa are not laid down, even in the mature<br />

basidium (Fig. 21.1a). The basidiospores are uninucleate<br />

and frequently germinate by repetition<br />

<strong>to</strong> produce uninucleate ballis<strong>to</strong>conidia. Eventually,<br />

hyphal germination occurs. Mating systems<br />

in the Cera<strong>to</strong>basidiales are not well unders<strong>to</strong>od,<br />

and several species may be homothallic.<br />

The basidium of the Tulasnellales differs in<br />

that the four epibasidia become separated from<br />

the metabasidium by septation after meiosis.<br />

Maturation of the four epibasidia may occur at<br />

different times (Fig. 21.1b).<br />

21.2.1 Rhizoc<strong>to</strong>nia in agriculture<br />

The most important taxon is R. (Moniliopsis) solani<br />

(teleomorph Thanatephorus cucumeris), which causes<br />

a wide array of soil-borne necrotrophic diseases<br />

especially of herbaceous plants including all<br />

kinds of vegetables, rice, turfgrasses, and less<br />

frequently also woody tree hosts (Adam, 1988;<br />

Agrios, 2005). The most common disease is<br />

damping off of seedlings in which the infected<br />

hypocotyl region becomes water-soaked and no<br />

longer provides structural integrity (see p. 95),<br />

leading <strong>to</strong> pre- or post-emergence death of<br />

the seedling. In older plants, infection may not<br />

immediately cover the entire circumference of<br />

the plant stem, so that cankers and girdling of the<br />

stem may result. Roots are also frequently<br />

infected, as are aerial plant organs in contact<br />

with the soil or exposed <strong>to</strong> watersplash from the<br />

soil surface. The fungus can survive in the soil<br />

for several years as small sclerotia about 1 3mm<br />

in diameter. These are often associated with<br />

the debris of host plants killed by the fungus.<br />

Sclerotia are occasionally seen as black scurf on<br />

the surface of pota<strong>to</strong> tubers because they are not<br />

easily removed, even by assiduous washing.<br />

Rhizoc<strong>to</strong>nia diseases are typically most severe<br />

at cool temperatures, presumably because the<br />

seedling stage of host plants is prolonged due<br />

<strong>to</strong> their slow growth. The infection process<br />

is similar <strong>to</strong> that of Gaeumannomyces graminis,<br />

i.e. hyphae branch and build up a cushion-like<br />

compound appressorium on the host plant<br />

surface, from which penetration is achieved.<br />

Fig 21.1 (a) Basidium of Thanatephorus cucumeris<br />

(Cera<strong>to</strong>basidiales). Note the four prominent<br />

epibasidia and the lack of septation within the<br />

basidium. (b) Basidium of Gloeotulasnella cystidiophora<br />

(Tulasnellales) showing various stages of maturity.<br />

The epibasidia are separated from the metabasidium<br />

by septation.Three epibasidia have already<br />

discharged their basidiospore, with one about <strong>to</strong><br />

produce it. (a) redrawn from Warcup and Talbot<br />

(1962) with permission from Elsevier, (b) redrawn<br />

from Wells and Bandoni (2001) with kind permission<br />

of Springer Science and Business Media.


596 HETEROBASIDIOMYCETES<br />

Simple appressoria may also be formed. Several<br />

lytic enzymes including cellulases, pectinases<br />

and cutinases are involved in colonizing and<br />

degrading plant tissue.<br />

Another pathogenic species is Rhizoc<strong>to</strong>nia<br />

cerealis (¼ Cera<strong>to</strong>rhiza cerealis, C. ramicola; teleomorph<br />

Cera<strong>to</strong>basidium cornigerum ¼ C. cereale),<br />

which causes sharp eyespot of cereals. As in<br />

R. solani, infection is favoured by cool conditions,<br />

and the disease is more common on winter<br />

cereals than spring-sown varieties. The fungus<br />

overwinters on stubble as mycelium and sclerotia,<br />

and infections give rise <strong>to</strong> spindle-shaped<br />

eyespot lesions on stems and leaves. These differ<br />

from true eyespot caused by Tapesia yallundae<br />

(p. 439) in being more sharply demarcated, with a<br />

reddish-brown margin enclosing an inner greyish<br />

region. Crop losses in cereals due <strong>to</strong> R. cerealis are<br />

not generally severe so that chemical control is<br />

not practised specifically against sharp eyespot<br />

(Parry, 1990). Rhizoc<strong>to</strong>nia cerealis also causes yellow<br />

or brown patches in swards of turfgrass, as well<br />

as root and foliar infections in a range of other<br />

crop plants (Kataria & Hoffman, 1988).<br />

Whereas chemical control of Rhizoc<strong>to</strong>nia spp.,<br />

like that of many other soil-borne pathogens, is<br />

difficult, biological control shows some promise.<br />

Several biocontrol organisms are effective under<br />

controlled labora<strong>to</strong>ry and greenhouse conditions,<br />

including species of Bacillus, Serratia and<br />

Strep<strong>to</strong>myces, as well as 2,4-diacetylphloroglucinolproducing<br />

Pseudomonas spp. (see p. 385). The high<br />

efficacy of Trichoderma spp. against Rhizoc<strong>to</strong>nia has<br />

been partially correlated with their secretion of<br />

cell wall-degrading enzymes, notably chitinases<br />

and b-(1,3)-glucanases (Innocenti et al., 2003;<br />

Markovich & Kononova, 2003). As in the case of<br />

the cereal take-all pathogen Gaeumannomyces<br />

graminis, certain soils can acquire the capacity<br />

<strong>to</strong> suppress Rhizoc<strong>to</strong>nia after several consecutive<br />

cultivation cycles with the same crop (Henis et al.,<br />

1979; Mazzola, 2002), and this has been attributed<br />

<strong>to</strong> the build-up of biocontrol organisms, especially<br />

Trichoderma and Pseudomonas spp.<br />

21.2.2 Rhizoc<strong>to</strong>nia and orchid mycorrhiza<br />

Stimulating accounts of this <strong>to</strong>pic have been<br />

written by Arditti (1992), Smith and Read (1997),<br />

Peterson et al. (1998) and Rasmussen (2002).<br />

All members of the plant family Orchidaceae<br />

(about 17 500 species) appear <strong>to</strong> be associated<br />

with mycorrhizal fungi at all stages of their life<br />

cycle in nature. In contrast <strong>to</strong> other types of<br />

mycorrhiza, there is a net flow of sugars from the<br />

fungal partner <strong>to</strong> the plant at least during the<br />

establishment of the orchid seedling. All colourless<br />

(non-pho<strong>to</strong>synthetic) orchids continue <strong>to</strong> rely<br />

on this external supply throughout their lives,<br />

and even green orchids do not seem <strong>to</strong> share<br />

their pho<strong>to</strong>synthetic products with the fungus<br />

(Alexander & Hadley, 1985). In orchid mycorrhiza,<br />

therefore, the plant parasitizes the fungus,<br />

and it is a curious fact that many of the fungi<br />

thus exploited are themselves serious plant<br />

pathogens, especially Rhizoc<strong>to</strong>nia spp. (Roberts,<br />

1999). Indeed, the very Rhizoc<strong>to</strong>nia strains isolated<br />

as pathogens of other plants (e.g. R. solani,<br />

R. cerealis) can support the germination of orchid<br />

seeds (Figs. 21.2b f). These as well as other<br />

species (e.g. R. goodyerae-repentis, R. repens) can<br />

also be isolated from mature orchid roots. Orchid<br />

mycorrhizal symbiosis therefore seems <strong>to</strong> be less<br />

specific than other forms of mycorrhiza<br />

(Masuhara et al., 1993).<br />

Orchid seeds are tiny and lack differentiated<br />

embryos or food reserves. In the absence of soluble<br />

external carbohydrates, they show only<br />

limited germination <strong>to</strong> form an intermediate<br />

stage called a pro<strong>to</strong>corm. This may emit a few<br />

epidermal hairs before growth stalls (Fig. 21.2b).<br />

Further development of the pro<strong>to</strong>corm (Fig. 21.2c)<br />

occurs only if a suitable soluble carbon source is<br />

added, or if a mycorrhizal fungus such as<br />

Rhizoc<strong>to</strong>nia is allowed <strong>to</strong> grow from a food base<br />

(e.g. starch or cellulose) <strong>to</strong> the pro<strong>to</strong>corm. Growth<br />

ensues even if the fungus is made <strong>to</strong> cross a<br />

barrier between the food base and the pro<strong>to</strong>corms,<br />

thereby demonstrating net carbon translocation<br />

(Smith, 1966). This experiment is easily<br />

set up in the labora<strong>to</strong>ry (Fig. 21.2d; Weber &<br />

Webster, 2001b). The main transport compound<br />

seems <strong>to</strong> be trehalose, and this may be hydrolysed<br />

<strong>to</strong> glucose and converted <strong>to</strong> sucrose by the plant<br />

(Smith, 1967; Smith & Read, 1997).<br />

Infection of the orchid pro<strong>to</strong>corm is initiated<br />

through the epidermal hairs (Fig. 21.2e) or<br />

through the suspensor tissue at the base of


CERATOBASIDIALES<br />

597<br />

Fig 21.2 Rhizoc<strong>to</strong>nia cerealis and its mycorrhiza with the heath spotted orchid (Dactylorhiza maculata ssp. erice<strong>to</strong>rum)inthe<br />

labora<strong>to</strong>ry. (a) Vegetative hypha showing typical branching and dolipore septa (arrowheads). (b) Seeds of D. maculata on agar after<br />

ten weeks without Rhizoc<strong>to</strong>nia. Pro<strong>to</strong>corms with a few epidermal hairs have formed. (c) Seeds after ten weeks but with R. cerealis<br />

spreading from a food base.The pro<strong>to</strong>corms have grown and are differentiating shoot tips (arrows). Same scale as (b). (d) The<br />

split-plate experiment. Rhizoc<strong>to</strong>nia cerealis has been inoculated on<strong>to</strong> a food base (tissue paper, <strong>to</strong>p half) which is separated from<br />

the orchid seeds by a partition.The fungus has overgrown this barrier, and the orchid pro<strong>to</strong>corms are using the translocated sugars<br />

derived from the degraded cellulose; 13 weeks after inoculation. (e) Penetration of an epidermal hair by R. cerealis. (f) Penetration of<br />

R. cerealis hyphae through an epidermal hair in<strong>to</strong> the cortex of a pro<strong>to</strong>corm where pelo<strong>to</strong>ns have formed. (b) and (d f) reprinted<br />

from Weber and Webster (2001b), with permission from Elsevier.<br />

the seedling. Penetration of the wall of a cell<br />

in the pro<strong>to</strong>corm cortex invaginates the plasmalemma<br />

and results in the formation of a pelo<strong>to</strong>n,<br />

i.e. a dense mass of coiled hyphae (Fig. 21.2f).<br />

Initially, each hypha is ensheathed by the host<br />

plasmalemma, which is called the perifungal<br />

membrane and is functionally modified from the<br />

plasmalemma of uninfected regions (Peterson &<br />

Currah, 1990; Peterson et al., 1996). An interfacial<br />

matrix of unknown composition is located<br />

between the perifungal membrane and the<br />

fungal cell wall. Within 24 h of formation, a<br />

pelo<strong>to</strong>n may begin <strong>to</strong> be degraded and is ultimately<br />

left behind as an amorphous clump<br />

of lysed hyphae surrounded by one continuous<br />

perifungal membrane (Hadley & Williamson,<br />

1971; Peterson & Currah, 1990). Any one<br />

orchid cell can become repeately re-infected


598 HETEROBASIDIOMYCETES<br />

(Uetake et al., 1992), and the same pro<strong>to</strong>corm,<br />

mature root or even individual cell can be<br />

colonized simultaneously by different fungi.<br />

The unstable nature of the orchid mycorrhiza<br />

is indicated by the quick and repeated cycle of<br />

pelo<strong>to</strong>n formation and degradation, and the<br />

several different outcomes of the orchid fungus<br />

interaction observed under labora<strong>to</strong>ry conditions<br />

(Fig. 21.2d). A balanced mycorrhizal symbiosis<br />

will develop only in a proportion of<br />

pro<strong>to</strong>corms, whereas other seeds of the same<br />

orchid species may be parasitized and killed by<br />

the fungus, or simply resist infection and stall in<br />

their development (Hadley, 1970; Smreciu &<br />

Currah, 1989; Beyrle et al., 1995). There is also<br />

vevidence of a succession of mycorrhizal fungi<br />

during the development of an orchid in nature,<br />

and the mycorrhizal fungi isolated from adult<br />

plants may not support pro<strong>to</strong>corm growth and<br />

vice versa (Xu & Mu, 1990; Zelmer et al., 1996).<br />

Mature orchids may be associated with<br />

Rhizoc<strong>to</strong>nia spp. and/or a range of other Basidiomycota,<br />

including saprotrophic (e.g. Mycena),<br />

necrotrophic (e.g. Armillaria; see p. 546) or<br />

ec<strong>to</strong>mycorrhizal species related <strong>to</strong> Russula and<br />

Thelephora (Rasmussen, 2002). Ec<strong>to</strong>mycorrhizal<br />

fungi transport carbohydrates from their tree<br />

host <strong>to</strong> the orchid, thus allowing green orchids<br />

<strong>to</strong> grow even in densely shaded woodland<br />

conditions (Taylor & Bruns, 1997; McKendrick<br />

et al., 2000; Bidar<strong>to</strong>ndo et al., 2004). Interestingly,<br />

these fungi form typical ec<strong>to</strong>mycorrhiza with<br />

their tree hosts but pelo<strong>to</strong>ns in infected orchid<br />

roots (Zelmer & Currah, 1995). The orchid therefore<br />

calls the shots in its symbiosis with<br />

basidiomycetes, and pelo<strong>to</strong>n formation and<br />

degradation is traditionally interpreted as a<br />

balanced defence reaction of the orchid against<br />

invasion attempts by the fungus.<br />

21.3 Dacrymycetales<br />

This order is characterized by forked (furcate)<br />

basidia (Figs. 21.4 and 21.5), which are found in<br />

all species except Dacrymyces unisporus. The fruit<br />

bodies are coloured yellow or orange due <strong>to</strong> the<br />

presence of a wide range of carotenoids (for<br />

references, see Gill & Steglich, 1987). Fruit bodies<br />

are gelatinous and show a striking diversity of<br />

forms as exemplified by the cushion-like basidiocarps<br />

of Dacrymyces stillatus (Fig. 21.3) and the<br />

clavarioid ones of Calocera viscosa (Plate 11g).<br />

Not much is known about the life cycle of<br />

the Dacrymycetales, but it is presumed that<br />

the usual basidiomycete pattern of alternating<br />

mono- and dikaryotic stages operates. There are<br />

no clamp connections. The dolipore-type septa<br />

Fig 21.3 Fruit bodies of Dacrymycetales. (a) Basidial<br />

cushions of Dacrymyces stillatus on rotting wood.<br />

(b) Basidiocarps of Calocera cornea.


DACRYMYCETALES<br />

599<br />

Fig 21.4 Dacrymyces stillatus. (a) Basidiospores showing germination by germ tubes or formation of conidia (bot<strong>to</strong>m). (b) Basidia.<br />

Note that the attached basidiospores are unicellular.They become three-septate on germination. (c) Arthrospores from a conidial<br />

pustule.<br />

are surrounded by parenthesomes without<br />

perforations (Wells, 1994). The probasidium<br />

arises from a dikaryotic hypha and is initially<br />

club-shaped. At this stage karyogamy occurs and<br />

is immediately followed by meiosis. Meanwhile<br />

the two epibasidia develop. Each of the two<br />

basidiospores seems <strong>to</strong> receive one nucleus, and<br />

the remaining two nuclei degenerate in the<br />

epibasidia (see Wells & Bandoni, 2001). Before<br />

the basidiospore germinates, it lays down one or<br />

more septa. Each spore segment can produce a<br />

haploid monokaryotic hypha or may give rise <strong>to</strong><br />

conidia which in turn germinate by means of<br />

monokaryotic germ tubes (Ingold, 1983b). It is<br />

unclear in many species how and where dikaryotization<br />

occurs, but it is probably by fusion<br />

of monokaryotic hyphae. Mating systems, where<br />

known, are bifac<strong>to</strong>rial (tetrapolar), i.e. with two<br />

mating type loci A and B.<br />

Members of the Dacrymycetales are saprotrophic<br />

on wood and cause brown-rots, although<br />

some lignin degradation has also been observed<br />

(Seifert, 1983; Worrall et al., 1997). The fruit<br />

bodies are common on decaying wood, including<br />

wood built in<strong>to</strong> outdoor structures such as park<br />

benches or fences. There are about 70 species,<br />

and the order is monophyletic (Oberwinkler,<br />

1993). Reid (1974) has given keys and descriptions<br />

of the common British and European species.<br />

The seminal features of the order have been<br />

summarized by Wells (1994) and Wells and<br />

Bandoni (2001).


600 HETEROBASIDIOMYCETES<br />

21.3.1 Dacrymyces<br />

Orange gelatinous cushions about 1 5mm in<br />

diameter, so common on damp rotting wood,<br />

are the fructifications of D. stillatus (Fig. 21.3a).<br />

Close inspection with a hand lens reveals that<br />

the fruit bodies are of two kinds: soft, bright<br />

orange, hemispherical cushions and firmer, pale<br />

yellow, flatter structures. The bright orange<br />

cushions are conidial pustules which consist of<br />

hyphae whose tips are branched and fragment<br />

in<strong>to</strong> numerous dikaryotic arthroconidia (Fig.<br />

21.4c). The cells are packed with oil globules<br />

containing carotenoids. Such conidia are readily<br />

dispersed by rainsplash and are obviously similar<br />

in function <strong>to</strong> the splash-dispersed conidia of<br />

Nectria cinnabarina (see Plate 5d). The yellow<br />

flatter structures are basidial cushions which<br />

are attached centrally <strong>to</strong> the woody substratum.<br />

The surface layer is composed of clusters of<br />

forked basidia (Fig. 21.4b) which arise from<br />

dikaryotic hyphae. Each basidium forms two<br />

haploid basidiospores. After discharge, a basidiospore<br />

undergoes nuclear division and<br />

septation <strong>to</strong> give four cells. Depending on environmental<br />

conditions, each cell germinates by<br />

means of a monokaryotic haploid germ tube<br />

or by a short conidiophore (denticle). Conidia<br />

may also arise on older hyphae. They germinate<br />

<strong>to</strong> give monokaryotic hyphae. Dikaryotization<br />

occurs when two compatible monokaryotic<br />

hyphae fuse. Mossebo and Amougou (2001) have<br />

shown that the parenthesome and dolipore<br />

complex dissolve in order <strong>to</strong> facilitate passage<br />

of nuclei through the septum in the course of<br />

dikaryotization. Dacrymyces stillatus is heterothallic<br />

with a bifac<strong>to</strong>rial (tetrapolar) mating system.<br />

Cells of the secondary mycelium are usually but<br />

not unfailingly dikaryotic (Mossebo, 1998).<br />

21.3.2 Calocera<br />

At first sight the ubiqui<strong>to</strong>us cylindrical orange<br />

outgrowths of C. viscosa from coniferous logs<br />

(Plate 11g), or the smaller C. cornea from hardwood<br />

logs (Fig. 21.3b), could be mistaken for<br />

species of Clavaria. However, the gelatinous<br />

texture and the characteristically forked basidia<br />

(Fig. 21.5) place them in the Dacrymycetales, and<br />

this placement has been confirmed by molecular<br />

phylogenetic studies (Weiss & Oberwinkler, 2001).<br />

Ingold (1983b) has carefully observed the fate<br />

Fig 21.5 Calocera viscosa. (a) T.S. through hymenium with furcate basidia at different stages of development. (b) Freshly discharged<br />

basidiospores which are aseptate.The clear area in each spore shows the displacement of cy<strong>to</strong>plasmic contents by the single<br />

nucleus. (c) Two 24-hour-old basidiospores on tap-water agar. Each spore has produced two conidiophores bearing microconidia<br />

on denticles. (d) Direct germination of a basidiospore has given rise <strong>to</strong> monokaryotic hyphae which are forming microconidia.<br />

The septum dividing the spore in<strong>to</strong> two is clearly visible. (b d) <strong>to</strong> same scale.


AURICULARIALES<br />

601<br />

of basidiospores in C. viscosa. Freshly discharged<br />

basidiospores are aseptate (Fig. 21.5b) but they<br />

soon develop one septum. Further development is<br />

by direct germination or formation of globose<br />

microconidia from basidiospore segments and<br />

from haploid monokaryotic hyphae (Figs. 21.5c,d),<br />

as described above for D. stillatus.<br />

21.4 Auriculariales<br />

The order Auriculariales has been subject <strong>to</strong><br />

numerous taxonomic rearrangements. As currently<br />

unders<strong>to</strong>od, its members produce both<br />

mono- and dikaryotic mycelia with dolipores<br />

and parenthesomes lacking perforations (see Fig.<br />

18.10b). Basidia are septate (Wells, 1994; Wells &<br />

Bandoni, 2001). The Auriculariales are distinguishable<br />

from the Tremellales (Section 21.5)<br />

which have yeast-like monokaryotic stages, and<br />

from the Cera<strong>to</strong>basidiales and Dacrymycetales<br />

which have aseptate basidia. Although taxonomic<br />

adjustments continue <strong>to</strong> be made, there is now<br />

little doubt that the order Auriculariales should<br />

contain both Auricularia with its transversely<br />

septate basidia, and Exidia and Pseudohydnum,<br />

which have basidia with longitudinal septa,<br />

the so-called tremelloid basidia (Weiss &<br />

Oberwinkler, 2001).<br />

There are great variations in fruit body size<br />

and shape. Clamp connections may be present<br />

or absent, depending on species. Most species<br />

have a bifac<strong>to</strong>rial mating system with multiple<br />

alleles at both loci (Wells, 1987, 1994; Wong &<br />

Wells, 1987; Wong, 1993). Depending on environmental<br />

conditions, basidiospores are typically<br />

able <strong>to</strong> germinate in several different ways,<br />

e.g. by repetition (ballis<strong>to</strong>spore formation), as<br />

hyphae, or by forming sickle-shaped (lunate)<br />

microconidia. Members of the Auriculariales<br />

are saprotrophic on wood, causing intensive<br />

white-rots (Worrall et al., 1997).<br />

21.4.1 Auricularia<br />

The Jew’s ear fungus A. auricula-judae forms<br />

rubbery, ear-shaped fruit-bodies on branches of<br />

elder (Sambucus) (Plate 11h) and is a weak<br />

pathogen, growing on the wood and pith of<br />

living branches and on dead wood. A wide range<br />

of other hosts has been reported, on which<br />

A. auricula-judae causes a rapid white-rot similar<br />

<strong>to</strong> that produced by members of the polyporoid<br />

clade (see Plate 10a; Worrall et al., 1997).<br />

A section through the flesh of a fruit body<br />

shows a hairy upper surface, a central gelatinous<br />

layer containing narrow clamped hyphae, and<br />

a broad hymenium on the lower side (Fig. 21.6a).<br />

Details of basidiocarp ana<strong>to</strong>my are useful in<br />

classification (Lowy, 1952). The fruit body can dry<br />

<strong>to</strong> a hard brittle mass, but on wetting it quickly<br />

absorbs moisture and discharges spores within<br />

a few hours. The basidia are cylindrical and<br />

become divided in<strong>to</strong> four cells by three transverse<br />

septa (Fig. 21.6b). Each cell of the basidium<br />

develops a long cylindrical epibasidium which<br />

extends <strong>to</strong> the surface of the hymenium and<br />

terminates in a conical sterigma bearing a<br />

basidiospore which is monokaryotic. At 20 mm<br />

or more in length, the Auricularia basidiospore<br />

is one of the largest objects propelled by the<br />

surface tension catapult mechanism (see Pringle<br />

et al., 2005), and ballis<strong>to</strong>spore discharge is easily<br />

observed with thin slices of basidiocarp material<br />

placed sideways on water agar (see Webster &<br />

Hard, 1998b). Auricularia auricula-judae is heterothallic<br />

with a bifac<strong>to</strong>rial (tetrapolar) mating<br />

system, and there are indications of multiple<br />

alleles (see Wong, 1993).<br />

Germination of the basidiospore can proceed<br />

in several different ways (Fig. 21.6c), and such a<br />

variability is common in the Auriculariales<br />

(Ingold, 1982a, 1984b). Washings from the hymenial<br />

surface of fruit bodies contain basidiospores<br />

undergoing repetitious germination by means<br />

of a sterigma which produces another ballis<strong>to</strong>spore.<br />

Ingold (1982a) interpreted this as a second<br />

chance for the spore <strong>to</strong> get away from the fruit<br />

body. Basidiospores alighting on a nutrient-poor<br />

surface such as tap water agar lay down three<br />

transverse septa, and each of the resulting four<br />

cells may emit one or more extensions (denticles)<br />

which produce a cluster of lunate microconidia.<br />

Alternatively or additionally, germination may<br />

occur directly by means of a germ tube, and<br />

this mode of germination is found especially<br />

on slightly richer media such as cornmeal agar.<br />

Septa are laid down, the first one bulging


602 HETEROBASIDIOMYCETES<br />

Fig 21.6 Auricularia auricula-judae.<br />

(a) Section of fruit body.The<br />

hymenium is on the lower side.<br />

(b) Squash preparation of the<br />

hymenium showing basidia. Note the<br />

transverse segmentation and the<br />

long epibasidia.The basidia are<br />

associated with branched hyphae.<br />

(c) Basidiospores.Two are<br />

ungerminated; one has developed<br />

a septum and is germinating directly<br />

by means of a germ tube; and<br />

two basidiospores have become<br />

three-septate and are producing<br />

lunate conidia from short<br />

conidiophores (denticles).<br />

(d) A basidiospore which had fallen<br />

back on<strong>to</strong> the hymenial surface and is<br />

germinating repetitiously <strong>to</strong> form a<br />

ballis<strong>to</strong>conidium. (b d) <strong>to</strong> same scale.<br />

backwards from the pro<strong>to</strong>plast-containing<br />

germ tube <strong>to</strong>wards the empty basidiospore.<br />

Such septa are interpreted as retraction septa<br />

(Ingold, 1982a). A richly branched monokaryotic<br />

mycelium of very fine hyphae develops, and after<br />

a while these hyphae form lateral or terminal<br />

denticles which produce clusters of lunate<br />

microconidia.<br />

The lunate conidia are a feature found in<br />

many species of Auriculariales, but their significance<br />

in the life cycle is uncertain. Like basidiospores<br />

and the repetitious ballis<strong>to</strong>conidia<br />

produced from them, they are capable of germination<br />

<strong>to</strong> form monokaryotic hyphae, but they<br />

might also function directly as spermatia. The<br />

putative life cycle of A. auricula-judae is shown in<br />

Fig. 21.7.<br />

Another common species is A. mesenterica,<br />

which forms thicker, hairy, fan-shaped fruit<br />

bodies on old stumps and logs of elm (Ulmus)<br />

and other trees. It, <strong>to</strong>o, causes active wood<br />

decay and may occasionally be weakly<br />

pathogenic.<br />

Auricularia as a cultivated mushroom<br />

Although devoid of any distinctive taste, the<br />

fruit bodies of Auricularia are highly nutritious<br />

and possess a chewy, rubbery texture which<br />

renders them attractive ingredients for Far<br />

Eastern soups and stir fries. The main species<br />

cultivated for food is A. polytricha (‘Mu-Erh’).<br />

The his<strong>to</strong>ry of cultivation dates back <strong>to</strong> AD 600<br />

in China (Cheng & Tu, 1978; Chang & Miles,<br />

2004), making Auricularia the first cultivated<br />

mushroom for which we have his<strong>to</strong>rical<br />

records. Some 465 000 t of fresh fruit bodies are<br />

currently produced per annum (Pegler, 2001).<br />

Very conveniently, the fruit bodies can be<br />

s<strong>to</strong>red dry for several months, and rehydrated<br />

when needed. Cultivation is traditionally performed<br />

by inoculating logs of suitable broadleaved<br />

trees with mycelial spawn. Infected logs<br />

can produce good crops for several years. The<br />

fungus is now also often cultivated in plastic<br />

bags filled with a mixture of sawdust and rice<br />

bran, allowing the fruit bodies <strong>to</strong> emerge<br />

through holes in the plastic.


AURICULARIALES<br />

603<br />

Fig 21.7 Life cycle of Auricularia auricula-judae. Depending on the substrate, a basidiospore has various options by which <strong>to</strong><br />

germinate; these are equivalent for both mating types, but for clarity we show them here for only one of them (white nuclei).<br />

Basidiospores falling on<strong>to</strong> the fruit body hymenium may germinate by repetition <strong>to</strong> form another ballis<strong>to</strong>spore. Depending on the<br />

nutrient status, basidiospores may germinate by formation of a monokaryotic mycelium or by producing lunate microconidia.The<br />

latter may also be produced by monokaryotic hyphae.Conjugation leads <strong>to</strong> the establishment of a dikaryotic mycelium which may<br />

form basidiocarps. Key events in the life cycle are plasmogamy (P), karyogamy (K) and meiosis (M). Haploid nuclei are drawn as<br />

empty or filled circles; the diploid nucleus is drawn larger and half-filled.<br />

21.4.2 Other members of the<br />

Auriculariales<br />

Exidia<br />

Exidia glandulosa, sometimes called ‘witches’<br />

butter’, forms black rubbery fructifications on<br />

decaying branches of various woody hosts, especially<br />

lime (Tilia) and oak (Quercus) (Fig. 21.9a).<br />

The hymenium is borne on the lower side of the<br />

fruit body and in this species is studded with<br />

small black warty outgrowths. The basidia are<br />

formed deep within the hymenium and produce<br />

long epibasidia. They are divided by longitudinal<br />

instead of transverse septa (Fig. 21.8).<br />

Basidiospores of E. glandulosa show germination<br />

patterns identical <strong>to</strong> those described above<br />

for A. auricula-judae, and Ingold (1982b) found<br />

that the only clear microscopic difference<br />

between these two species is the arrangement<br />

of septa in the basidium. Phylogenetic studies<br />

have confirmed the close relationship between<br />

Auricularia and Exidia (Weiss & Oberwinkler, 2001).<br />

Pseudohydnum gelatinosum<br />

This species grows on dead stumps and branches<br />

of coniferous trees. The fruit body is jelly-like in<br />

consistency and has a short eccentric stalk.<br />

The hymenium is on the lower side of the<br />

pileus and is arranged in<strong>to</strong> numerous conical<br />

teeth (Fig. 21.9b) resembling those of Hydnum.<br />

The basidia are similar <strong>to</strong> those of Exidia in<br />

being longitudinally septate. As in Exidia and<br />

Auricularia, basidiospores failing <strong>to</strong> escape from


604 HETEROBASIDIOMYCETES<br />

the hymenium may germinate by repetition or,<br />

in other situations, by means of a germ tube, but<br />

lunate or other microconidia have not been<br />

observed (Ingold, 1985).<br />

21.5 Tremellales<br />

Fig 21.8 Exidiaglandulosa.Sectionofhymeniumshowing<br />

longitudinally divided basidia with long epibasidia extending <strong>to</strong><br />

the surface.<br />

Formerly unders<strong>to</strong>od as a broad taxon which<br />

included genera such as Exidia and Pseudohydnum<br />

(see p. 603), the order Tremellales is now<br />

restricted <strong>to</strong> fungi which possess a yeast-like<br />

haploid state, basidia divided by longitudinal<br />

septa (tremelloid basidia), and dikaryotic hyphae<br />

with a dolipore septum and a parenthesome<br />

which is sacculate, i.e. invaginated <strong>to</strong>wards the<br />

septal pore (Fig. 21.10a; Berbee & Wells, 1988).<br />

Dikaryotic hyphae usually have clamp connections.<br />

The monokaryotic yeast state resembles<br />

heterobasidiomycete yeasts such as Filobasidiella<br />

neoformans, Phaffia rhodozyma and species of<br />

Bullera and Cryp<strong>to</strong>coccus, all of which belong <strong>to</strong><br />

the Tremellales and related orders within the<br />

Tremellomycetidae (Fell et al., 2001). Heterobasidiomycetes<br />

growing predominantly in the<br />

yeast state are discussed in Chapter 24 (p. 660).<br />

In the life cycle of the filamen<strong>to</strong>us Tremellales<br />

considered here, the dikaryotic condition is the<br />

dominant phase and is re-established by<br />

Fig 21.9 Fruit bodies of Auriculariales. (a) Exidiaglandulosa.Fruit bodies on lime (Tilia).The hymenial surface bears black warts and<br />

is on one face of the fruit body. (b) Pseudohydnum gelatinosum.Fruit bodies seen from above (right) and below (left).The hymenium<br />

covers the surface of the spines.


TREMELLALES<br />

605<br />

conjugation of compatible yeast cells. The mating<br />

system has been described as ‘modified tetrapolar’<br />

(Bandoni, 1963) because there are only two<br />

alleles at the A locus but multiple alleles at B.<br />

This system is uncommon in the Basidiomycota,<br />

which usually have multiple alleles at both loci,<br />

aiding outbreeding (see p. 507). The modified<br />

tetrapolar system of Tremella is, however, also<br />

found in Ustilago maydis, and the designation<br />

of loci is equivalent (p. 643). Thus, the A locus<br />

controls conjugation and the B locus growth of<br />

the resulting dikaryon. Both A alleles of Tremella<br />

encode peptide-type hormones (Sakagami et al.,<br />

1981; Ishibashi et al., 1984) and recep<strong>to</strong>rs for<br />

the hormone of the opposite mating type. In<br />

T. mesenterica, these hormones are linear peptides<br />

called tremerogen A-10 (12 amino acids) and<br />

tremerogen a-13 with 13 amino acids. Both<br />

peptides are derivatized with a farnesyl unit.<br />

Mating occurs by formation of conjugation tubes<br />

which requires the presence of the hormone<br />

of the opposite mating type; each hormone<br />

is produced constitutively, irrespective of the<br />

presence or absence of a compatible mating<br />

partner (Bandoni, 1965). Yeast cells with opposite<br />

alleles at A but like B alleles will conjugate but<br />

fail <strong>to</strong> initiate dikaryotic hyphal growth.<br />

The fruit bodies of Tremellales are usually<br />

formed on wood, often in association with those<br />

of other fungi (Asco- and Basidiomycota) or<br />

with lichen thalli, which may be parasitized<br />

(Diederich, 1996; Chen, 1998). Parasitism is by<br />

intimate hyphal contact via haus<strong>to</strong>rial branches<br />

(Figs. 21.10b,c; see below). The fruit bodies of<br />

Tremellales are highly variable in size, ranging<br />

from a limited hymenium on the mycelium of<br />

putative hosts <strong>to</strong> large structures (several centimetres)<br />

surrounding host basidiocarps or growing<br />

near them. Although often considered as<br />

saprotrophs, no special capacity <strong>to</strong> degrade wood<br />

seems <strong>to</strong> have been recorded.<br />

Accounts of the Tremellales have been written<br />

by Bandoni (1987), Bandoni and Boekhout<br />

(1998) and Chen (1998).<br />

21.5.1 Tremella<br />

This is a large genus of some 80 species. Detailed<br />

descriptions of representatives of all species<br />

groups have been given by Chen (1998). One of<br />

the commonest and most thoroughly examined<br />

species is T. mesenterica, whose yellow or orange<br />

gelatinous fruit bodies are readily seen on<br />

various woody hosts such as oak, willow, gorse<br />

and beech (Plate 11i). Variations in the intensity<br />

of fruit body coloration could be due <strong>to</strong> a<br />

stimulation of carotenoid synthesis by high<br />

light intensity because fruit bodies exposed <strong>to</strong><br />

sunlight are often more deeply coloured than<br />

those in the shade (Wong et al., 1985). The fruit<br />

bodies of T. mesenterica are usually associated<br />

with those of Peniophora in the field, and<br />

Zugmaier et al. (1994) have shown that hyphae<br />

of Peniophora spp. are parasitized in vivo and in<br />

vitro. Several Tremella spp. are more obviously<br />

mycoparasitic than T. mesenterica. For instance,<br />

T. globospora produces its fruit bodies within the<br />

perithecia of Diaporthe, and T. encephala overgrows<br />

the fructifications of its host, Stereum, which<br />

remain as a firm core in the Tremella basidiocarp.<br />

There are also several other genera within the<br />

Tremellales which parasitize other fungi (Bauer,<br />

2004; Figs. 21.10b,c).<br />

Parasitism is mediated by modified hyphae<br />

which Olive (1947) called ‘haus<strong>to</strong>rial branches’.<br />

They consist of a swollen binucleate hyphal<br />

segment, delimited at its base by a clamp<br />

connection which puts forward one or several<br />

long thin filaments (Fig. 21.11a). The association<br />

of the swollen segment with the term ‘haus<strong>to</strong>rium’<br />

is unfortunate because the haus<strong>to</strong>rial<br />

branch is formed outside the host cell. Where<br />

‘haus<strong>to</strong>rial filaments’ of Tremella or related<br />

fungi contact the hypha of a suitable host, the<br />

host wall is dissolved and a micropore is formed<br />

which establishes direct cy<strong>to</strong>plasmic contact<br />

between the filament and the host, apparently<br />

by fusion of the two plasma membranes<br />

(Figs. 21.10b,c).<br />

The life cycle of T. mesenterica is complex and<br />

not yet fully unders<strong>to</strong>od (Fig. 21.12). Haploid<br />

basidiospores are discharged from long epibasidia<br />

by the surface tension catapult mechanism.<br />

Basidiospores failing <strong>to</strong> escape from the hymenium<br />

can germinate by repetition <strong>to</strong> form<br />

ballis<strong>to</strong>conidia (Ingold, 1982b). When landing on<br />

a suitable substrate, the basidiospore germinates<br />

by forming several buds which remain attached


606 HETEROBASIDIOMYCETES<br />

Fig 21.10 Ultrastructure of Tremellales. (a) Cupulate parenthesome and dolipore septum of T. globospora.(b,c)Tipofthe<br />

haus<strong>to</strong>rial filament of Trimorphomyces papilionaceus (<strong>to</strong>p) which has made contact with a hypha of its host, Arthrinium<br />

sphaerospermum (bot<strong>to</strong>m). Membrane continuity has been established in the micropore region (c).Original prints kindly provided<br />

by M.L.Berbee (a) and R.Bauer (b,c). a reprinted from Berbee and Wells (1988), with permission from Mycologia. ßThe Mycological<br />

Society of America.<br />

Fig 21.11 Tremella mesenterica. (a) Haus<strong>to</strong>rial branch.<br />

The swollen dikaryotic branch arises from a clamp<br />

connection and emits filaments which may contact<br />

host hyphae. (b) Dikaryotic clamped hypha from the<br />

hymenium of a fruit body. One branch has produced<br />

a basidium whereas the other has formed dikaryotic<br />

conidia. Redrawn from Wells and Bandoni (2001) with<br />

kind permission of Springer Science and Business<br />

Media.<br />

<strong>to</strong> the basidiospore. These buds act as conidiogenous<br />

cells by producing numerous minute<br />

blas<strong>to</strong>conidia, which in turn germinate by swelling<br />

and budding <strong>to</strong> give rise <strong>to</strong> the haploid yeast<br />

state. Fusion between compatible yeast cells<br />

re-establishes the dikaryotic mycelial phase.<br />

In addition <strong>to</strong> producing basidia, hyphae in the<br />

fruit body may also form a dikaryotic conidial


TREMELLALES<br />

607<br />

Fig 21.12 Life cycle of Tremellamesenterica.The monokaryotic part of the cycle is shown only for one of the two mating types<br />

(white nuclei).Basidiospores failing <strong>to</strong> leave the fruitbody maygain a second chance byrepetitious germination <strong>to</strong> form a<br />

ballis<strong>to</strong>conidium.On other substrates, basidiospores germinate by way ofblas<strong>to</strong>conidia which give rise <strong>to</strong>yeastcells.Conjugation<br />

of two compatibleyeastcells givesrise <strong>to</strong> the dikaryotic mycelial stagewhich forms fruitbodiesproducing dikaryotic conidia, haploid<br />

yeastcells by de-dikaryotization (not shown), andhaploidbasidiosporesby meiosis.Key eventsin thelife cycle areplasmogamy (P),<br />

karyogamy (K) andmeiosis (M).Haploid nuclei are drawn as empty or filled circles; the diploid nucleus is drawn larger andhalf-filled.<br />

state (Fig. 21.11b) with an uncertain role in the life<br />

cycle. Additionally, de-dikaryotization <strong>to</strong> give<br />

yeast cells has been observed in the fruit bodies<br />

of T. mesenterica.<br />

Another common species is T. foliacea (sometimes<br />

synonymized with T. frondosa) with its fleshcoloured<br />

<strong>to</strong> pale brown, lobed or con<strong>to</strong>rted fruit<br />

bodies on oak and beech stumps. In this species,<br />

basidiospores germinate on suitable substrates<br />

by giving rise directly <strong>to</strong> budding yeast cells<br />

(Fig. 21.13). This has also been observed in<br />

T. encephala (Ingold, 1985). There are no dikaryotic<br />

conidia in these two species. Most Tremella spp. are<br />

heterothallic with a modified tetrapolar mating<br />

system, but in T. fuciformis both homothallic<br />

and heterothallic strains have been observed<br />

(Fox & Wong, 1990).<br />

Cultivation of Tremella<br />

The ‘silver ear’ fungus, T. fuciformis, has been<br />

cultivated in China on wood and sawdust for<br />

about 200 years (Chang & Miles, 2004). Although<br />

almost always associated with Hypoxylon spp.<br />

in nature (Chen, 1998), mycoparasitism does<br />

not seem <strong>to</strong> be obligate because cultivation is<br />

possible in monoculture. However, yields are<br />

greatly stimulated in the presence of a ‘friend of<br />

the mycelium’, i.e. the substrate is co-inoculated<br />

with T. fuciformis and Hypoxylon archeri or another<br />

suitable host species (Chang & Miles, 2004).


608 HETEROBASIDIOMYCETES<br />

Fig 21.13 Tremella frondosa. (a) Basidia showing epibasidia with various stages of spore development. (b) Freshly discharged<br />

basidiospore. (c) Basidiospore germinating on malt extract agar by budding <strong>to</strong> form yeast-like cells which also undergo budding.<br />

(d) Basidiospores germinating in water by repetition, i.e. by producing ballis<strong>to</strong>spores. (e) Immature basidia seen from above, showing<br />

the division of the basidium in<strong>to</strong> four cells by longitudinal septa.<br />

The fruit bodies of T. fuciformis are consumed in a<br />

stewed form, especially as a dessert, and this<br />

species has traditionally also been viewed as a<br />

medicinal mushroom in the Far East. Lifeprolonging,<br />

vitalizing and anti-cancer properties<br />

have been ascribed <strong>to</strong> T. fuciformis, and these<br />

effects seem <strong>to</strong> be due mainly <strong>to</strong> a stimulation of<br />

the immune system (Wasser, 2002). The biochemical<br />

basis is thought <strong>to</strong> be due <strong>to</strong> the production<br />

of exopolysaccharides by this and other Tremella<br />

spp., including T. mesenterica (Reshetnikov et al.,<br />

2000). However, it is as yet unclear how such<br />

polysaccharides are assimilated by the human<br />

digestive system, and how they interact with<br />

human cells. Many Homobasidiomycetes produce<br />

exopolysaccharides that are said <strong>to</strong> possess<br />

similar medicinal properties, and these molecules<br />

are mainly b-(1,3)-glucans substituted with<br />

various sugar moieties. In contrast, in Tremella<br />

the biologically active molecules have a b-(1,3)-<br />

mannan backbone substituted with xylose and<br />

glucuronic acid (de Baets & Vandamme, 2001;<br />

Vinogradov et al., 2004). A multitude of such<br />

mannans differing mainly in their substitution<br />

pattern is produced by the fruit bodies on<br />

solid substrata and by yeast cells in liquid<br />

culture. Yeast cells are encapsulated by these<br />

polysaccharides, as revealed by staining with<br />

Indian ink (see Fig. 24.1b). Structurally similar<br />

xyloglucuronomannans make up the capsule<br />

of the human pathogen Cryp<strong>to</strong>coccus neoformans<br />

(see p. 661).


22<br />

Urediniomycetes: Uredinales (rust fungi)<br />

22.1 Urediniomycetes<br />

Following extensive re-arrangements, the class<br />

Urediniomycetes (about 8000 species) is now<br />

considered <strong>to</strong> be monophyletic, although the<br />

naming of orders and families is still proving<br />

difficult (Swann & Taylor, 1995; Kirk et al., 2001;<br />

Swann et al., 2001). The order Uredinales (rust<br />

fungi) is by far the largest (about 7000 species)<br />

and the most important. The order Microbotryales,<br />

although taxonomically part of the Urediniomycetes,<br />

is a group of fungi causing smut<br />

diseases and will be discussed in Chapter 23.<br />

Many Urediniomycetes belonging <strong>to</strong> several<br />

orders occur predominantly in the yeast state.<br />

An important group, the Sporidiales, contains<br />

the red yeasts Sporidiobolus and Rhodosporidium<br />

(anamorphs Sporobolomyces and Rhodo<strong>to</strong>rula, respectively),<br />

and this order is considered in more<br />

detail on pp. 666 670.<br />

General information on the groups included<br />

in the Urediniomycetes has been given by Swann<br />

et al. (2001). They have defined Urediniomycetes<br />

as fungi in which the processes of karyogamy<br />

and meiosis occur in distinct parts of the basidium,<br />

i.e. the probasidium and metabasidium,<br />

respectively. The metabasidium is typically transversely<br />

septate, with basidiospores produced<br />

laterally (see Fig. 22.2d). Another useful character<br />

is the structure of septa viewed by<br />

transmission electron microscopy. Urediniomycete<br />

septa are simple with a single pore which<br />

may be open or plugged, but they typically<br />

lack the dolipore arrangement found in other<br />

basidiomycetes (see Fig. 18.9). Clamp connections<br />

are absent.<br />

22.2 Uredinales: the rust fungi<br />

Rust fungi (Uredinales) are a fascinating group of<br />

organisms. The life cycle of a typical rust species<br />

is among the most complex found anywhere in<br />

nature, consisting of five different spore stages<br />

on two plant hosts which are taxonomically<br />

entirely unrelated <strong>to</strong> each other. These pathogens<br />

infect most groups of vascular plants,<br />

including Pteridophytes (ferns), Gymnosperms,<br />

and Angiosperms (both monocots and dicots).<br />

Numerous fundamental questions about rust<br />

fungi remain <strong>to</strong> be answered, e.g. how a biotrophic<br />

organism manages <strong>to</strong> infect and parasitize<br />

two unrelated hosts using different<br />

mechanisms on either; how the five spore<br />

stages with their numerous different dispersal<br />

mechanisms could have evolved; how easily one<br />

or more of them can become aborted in derived<br />

(reduced) life cycles; how rust fungi survive in<br />

situations where one of their two hosts is<br />

unavailable; and how quickly new rust species<br />

or races spread <strong>to</strong> new habitats and then come <strong>to</strong><br />

an equilibrium with their host plants. A pro<strong>to</strong>col<br />

<strong>to</strong> generate stable transformants of rust fungi<br />

would greatly facilitate experimental work on<br />

them, but unfortunately this is not yet available.<br />

The species concept in rust fungi is also<br />

challenging. Morphological species are readily<br />

recognized, but these can show considerable


610 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

genetic adaptability by fragmenting in<strong>to</strong> several<br />

forms which infect non-overlapping spectra of<br />

host species and thereby become reproductively<br />

isolated. Because of the identity of the host<br />

family as an additional taxonomic feature, species<br />

of rust fungi are relatively easily grouped<br />

in<strong>to</strong> genera which are often monophyletic. At the<br />

higher taxonomic level, two large groups can be<br />

resolved weakly by phylogenetic analyses (Maier<br />

et al., 2003; Wingfield et al., 2004). They coincide<br />

with the Pucciniaceae and Melampsoraceae of<br />

earlier concepts (Dietel, 1928) in which these two<br />

families were distinguished by their teliospore<br />

or probasidium being stalked (Pucciniaceae) or<br />

unstalked (Melampsoraceae). Another point of<br />

distinction is that a particular spore-producing<br />

structure, the aecial stage (see below), is produced<br />

on angiosperm hosts by Pucciniaceae but<br />

on gymnosperms by Melampsoraceae (Wingfield<br />

et al., 2004). Dietel’s two broad groups are usually<br />

split up in<strong>to</strong> about 13 14 smaller families<br />

(Kirk et al., 2001; Cummins & Hiratsuka, 2003),<br />

but since the boundary lines are still being<br />

re-drawn from time <strong>to</strong> time, we prefer <strong>to</strong> adhere<br />

<strong>to</strong> the original concept for the purposes of this<br />

book.<br />

The popular name ‘rust fungi’ refers <strong>to</strong> the<br />

reddish-brown colour of some of the spores<br />

which are produced in dense pustules on crop<br />

plants, giving them a ‘rusted’ appearance. This<br />

is especially true of the rusts on cereals, which<br />

have probably caused crop losses since the beginning<br />

of agriculture. Archaeological excavations<br />

have uncovered rusted cereal remains dating<br />

back <strong>to</strong> the Bronze Age (Kislev, 1982), and it is<br />

well known that the ancient Romans held a<br />

special festival, the Robigalia, <strong>to</strong> appease their<br />

rust gods. This <strong>to</strong>ok place on 25 April, i.e. at a<br />

time when the crop was particularly vulnerable<br />

<strong>to</strong> attack (Large, 1940). The Robigalia may be the<br />

origin of Rogation Sunday, a day of blessing of<br />

the crops which is still observed by some<br />

Christian churches every year in late April<br />

(Schuman, 1991).<br />

Rusts can also cause serious economic damage<br />

on non-cereal crops, and examples are given in<br />

later sections of this chapter. Because of the<br />

immense economic importance of rust fungi,<br />

an enormous body of literature has been written,<br />

and the wheat Puccinia graminis system is probably<br />

the most thoroughly examined of all<br />

host pathogen interactions involving fungi.<br />

None the less, no substantial integrated treatment<br />

of the rust fungi has appeared in the past<br />

two decades, the two-volume set on cereal rusts<br />

(Bushnell & Roelfs, 1984; Roelfs & Bushnell, 1985)<br />

having been the last major effort. However, good<br />

keys and species descriptions are available, e.g.<br />

in Grove (1913), Gäumann (1959), Wilson and<br />

Henderson (1966) and Cummins and Hiratsuka<br />

(2003).<br />

22.2.1 The basic life cycle of rusts<br />

The classical example of a rust fungus, Puccinia<br />

graminis, is the cause of black stem rust on wheat<br />

and other cereals (Fig. 22.1), which is described<br />

more fully in Section 22.3. The life cycle of rusts<br />

is homologous with that of the Homobasidiomycetes<br />

(Fig. 18.4) in being divided in<strong>to</strong> stages of<br />

primary (homo- and mono-karyotic) and secondary<br />

(hetero- and di-karyotic) mycelium. The<br />

heterokaryotic phase is the main period in the<br />

life cycle of rusts and it is the only one in which<br />

some rusts can survive indefinitely under suitable<br />

conditions. The host plant species on which<br />

this stage is produced is therefore termed the<br />

principal host, with that bearing the homokaryotic<br />

mycelium called the alternate host.<br />

On leaves of the principal host, P. graminis<br />

produces urediniospores in pustules called<br />

uredinia, and they rapidly spread the infection<br />

because they are capable of re-infecting the same<br />

host species. Urediniospores are produced on<br />

stalks from which they break off at maturity,<br />

being released and distributed passively by wind.<br />

Urediniospores of rust fungi have a relatively<br />

thick wall which is often pigmented and typically<br />

spiny. They are capable of surviving airborne<br />

for several weeks or months, which accounts for<br />

their long-distance dispersal sometimes over<br />

hundreds or even thousands of miles. Each<br />

urediniospore is binucleate, with the two nuclei<br />

being of opposite mating type. In temperate<br />

climates, uredinia of many rusts are gradually<br />

replaced by telia in autumn, especially on leaf<br />

sheaths and stems. Teliospores (¼ probasidia)<br />

of Puccinia are two-celled and thick-walled.


UREDINALES: THE RUST FUNGI<br />

611<br />

Fig 22.1 The life cycle of Puccinia graminis, with its heterokaryotic phase on cereals and the homokaryotic stage on barberry<br />

(Berberis).The different ways in which the five spore stages are released and dispersed are indicated.Teliospores overwinter after<br />

nuclear fusion, i.e. as diploid cells. During basidiosporogenesis, meiosis is followed by a mi<strong>to</strong>tic division so that each basidiospore is<br />

a homokaryon containing two nuclei of the same mating type.Open and closed circles represent haploid nuclei of opposite mating<br />

type; diploid nuclei are larger and half-filled. Key events in the life cycle are plasmogamy (P), karyogamy (K) and meiosis (M).<br />

Each cell contains two haploid nuclei of opposite<br />

mating-type which fuse <strong>to</strong> form a diploid nucleus<br />

(see Fig. 22.2c). The teliospore overwinters in the<br />

diploid state, firmly attached <strong>to</strong> the plant tissue<br />

on which it was produced. In spring, meiosis<br />

occurs while each teliospore cell emits a germ<br />

tube which becomes transversely septate and<br />

forms the promycelium or metabasidium. Each<br />

compartment of the metabasidium produces one<br />

basidiospore which initially contains one haploid<br />

nucleus. This commonly divides so that the<br />

mature basidiospore often contains two genetically<br />

identical haploid nuclei and is thus a<br />

dikaryotic homokaryon (Anikster, 1983). Basidiospores<br />

are actively liberated by the surface<br />

tension catapult mechanism involving Buller’s<br />

drop as described before for other basidiomycetes<br />

(Section 18.5).<br />

The basidiospores of P. graminis are unable <strong>to</strong><br />

infect the cereal host but will infect the alternate<br />

host, barberry (Berberis vulgaris). The resulting<br />

primary mycelium is haploid, homo- and monokaryotic.<br />

At the upper (adaxial) surface of the<br />

host leaf, the primary mycelium forms a flaskshaped<br />

fructification known as a spermogonium.<br />

Within the mesophyll, knots of hyphae<br />

form a pro<strong>to</strong>-aecium but do not develop further<br />

at this stage. Minute uninucleate spermatia<br />

are produced from annellide-like structures<br />

(Littlefield & Heath, 1979) within the main body<br />

of the spermogonium, and they aggregate within<br />

a rim of hairs (periphyses) at the spermogonial


612 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

Fig 22.2 Examples of spore stages in rust fungi. (a) Uredinium of Puccinia obscura on its principal host, the wood-rush<br />

Luzula campestris. Note the spiny urediniospores. (b) Development of a telium within an old uredinium of P. obscura.Some<br />

urediniospores are still seen on <strong>to</strong>p of the smooth-walled teliospores pushing through. (c) Karyogamy in a teliospore of Puccinia<br />

lagenophorae, a demicyclic rust on Seneciovulgaris.The terminal cell still contains two paired nuclei (arrowheads) whereas in the basal<br />

cell the two nuclei have already fused <strong>to</strong> give one larger (diploid) nucleus. (d) Production of a promycelium (metabasidium) by Puccinia<br />

graminis.Four sterigmata have formed, each about <strong>to</strong> produce one basidiospore. (e) Spermogonium of P. obscura on its<br />

alternate host, Bellis perennis. Note the flask-shaped body and the ring of stiff periphyses which protrude beyond the upper host<br />

epidermis. (f) Aecia of P. lagenophorae on Senecio vulgaris.<br />

ostiole in a sugary liquid which attracts insects.<br />

These carry spermatia around, and fertilization<br />

occurs if a spermatium makes contact with<br />

a modified periphysis called receptive hypha<br />

of a spermogonium of opposite mating type.<br />

Spermatia are incapable of independent<br />

germination <strong>to</strong> establish new infections but<br />

perform solely a sexual role in fertilization.<br />

Most rust fungi are heterothallic with a bipolar<br />

mating system, although Narisawa et al. (1994)<br />

have claimed that P. coronata has a tetrapolar<br />

mating system with two mating type loci.


UREDINALES: THE RUST FUNGI<br />

613<br />

Homothallic species are also known, and these<br />

generally do not produce spermogonia. A generalized<br />

account of rust life cycles has been given<br />

by Buller (1950).<br />

Following fertilization, the nucleus from<br />

the donor spermatium divides repeatedly and<br />

migrates down the receptive hypha in<strong>to</strong> the<br />

receiving primary mycelium which thereby<br />

undergoes dikaryotization. When compatible<br />

nuclei reach the pro<strong>to</strong>-aecium, this becomes<br />

converted in<strong>to</strong> an aecium which breaches the<br />

lower epidermis. Heterokaryotic aeciospores<br />

develop and are released by a sudden roundingoff<br />

of the flattened wall separating adjacent<br />

spores, thus flicking the spores in<strong>to</strong> the air.<br />

Aeciospores are relatively thin-walled and bear<br />

warty rather than spiny surface ornamentations.<br />

They are often brightly coloured due <strong>to</strong> the<br />

abundance of carotenoids accumulating in lipid<br />

droplets (Plate 12a). Aeciospores of P. graminis<br />

are unable <strong>to</strong> re-infect Berberis, but are infective<br />

on the principal grass host. The resulting secondary<br />

mycelium produces urediniospores, thus<br />

completing the life cycle.<br />

Examples of the different spore types and<br />

the pustules (sori) producing them are shown in<br />

Fig. 22.2. Spores and sori have been given various<br />

names, and we follow the naming used in the<br />

Dictionary of <strong>Fungi</strong> (Kirk et al., 2001). It is cus<strong>to</strong>mary<br />

<strong>to</strong> assign Roman numerals 0, I, II, III or IV<br />

<strong>to</strong> the distinct spore types, and these numbers<br />

provide a convenient shorthand for describing<br />

the range of spores found in a given rust.<br />

The most important terms are summarized in<br />

Table 22.1.<br />

The various spore stages of rust fungi differ<br />

greatly in their length of survival. Although<br />

details are dependent on species and conditions<br />

of s<strong>to</strong>rage (temperature, state of hydration,<br />

light), it may be generalized that the maximum<br />

survival period in the field is in the order<br />

of days (spermatia and basidiospores), weeks<br />

(aeciospores), a few months (urediniospores)<br />

and several months <strong>to</strong> more than a year<br />

(teliospores).<br />

22.2.2 Derived life cycles and Tranzschel’s<br />

Law<br />

Rust fungi which must alternate between two<br />

different host plants in order <strong>to</strong> complete their<br />

life cycle are called heteroecious. In contrast,<br />

au<strong>to</strong>ecious species are confined <strong>to</strong> one host<br />

plant. Species whose life cycle contains all five<br />

possible spore stages are called macrocyclic. In<br />

this terminology, P. graminis is a macrocyclic<br />

heteroecious rust whereas P. menthae, which<br />

produces all five spore stages on one host<br />

(mint), is macrocyclic but au<strong>to</strong>ecious. Many<br />

rusts have derived life cycles in which one or<br />

more spore stages have been omitted. One<br />

common variation is the absence of uredinia in<br />

heterocyclic rusts. For instance, Gymnosporangium<br />

fuscum (see p. 629) produces spermogonia and<br />

aecia on the leaves of pear trees and has Juniperus<br />

spp. as its principal host for production of telia.<br />

Such rusts which lack uredinia are called<br />

demicyclic or -opsis forms.<br />

An au<strong>to</strong>ecious version of the demicyclic theme<br />

is presented by P. lagenophorae which produces<br />

spermogonia, aecia and telia on Senecio spp.<br />

(see Figs. 22.2c f). In functional terms, the aecia<br />

could be considered uredinia because the aeciospores<br />

infect the same host species on which they<br />

were produced, and because in old aecia the<br />

Table 22.1. Generally accepted terminology of the sori and spore states of rust fungi.<br />

0 I II III IV<br />

Sorus Spermogonium Aecium Uredinium Telium Basidium<br />

Pycnium Aecidiosorus Uredosorus Teleu<strong>to</strong>sorus Metabasidium<br />

Aecidium Uredium Promycelium<br />

Spore Spermatium Aeciospore Urediniospore Teliospore Basidiospore<br />

Pycniospore Aecidiospore Uredospore Teleu<strong>to</strong>spore Sporidium<br />

Urediospore


614 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

aeciospores can be displaced by teliospores. Such<br />

aecia are sometimes termed aecidioid uredinia.<br />

Another common variation is the microcyclic<br />

one in which both aecia and uredinia are<br />

lacking. Such rusts are, of course, au<strong>to</strong>ecious.<br />

Many microcyclic rusts do not undergo sexual<br />

reproduction in the sense of sexual recombination,<br />

i.e. spermogonia are absent and the only<br />

reproductive unit is the teliospore with or<br />

without basidiospores. Such species are either<br />

homothallic or asexual. The wide array of<br />

possible nuclear cycles has been summarized by<br />

Ono (2002).<br />

An example of a microcyclic rust fungus is<br />

P. mesnieriana, which produces telia on Rhamnus<br />

catharticus (buckthorn). The teliospores germinate<br />

<strong>to</strong> produce metabasidia, but only two basidiospores<br />

are formed per basidium (Anikster & Wahl,<br />

1985). This species is homothallic because a single<br />

basidiospore can infect R. catharticus, giving rise <strong>to</strong><br />

a telium-producing infection. The teliospores are<br />

highly characteristic because they carry a ‘crown’<br />

of spiny outgrowths. Very similar teliospores (see<br />

Fig. 22.13a) are produced by the macrocyclic<br />

‘crown rust’, Puccinia coronata, which also colonizes<br />

Rhamnus and related plants as alternate<br />

hosts, but infects grasses and cereals, especially<br />

oat (Avena), as the principal (i.e. uredinial and<br />

telial) host. Working in the late nineteenth and<br />

early twentieth century when the elucidation of<br />

rust life cycles was very much in vogue, the<br />

Russian mycologist Vladimir Tranzschel generalized<br />

that an unknown aecial stage of a macrocyclic<br />

uredinial/telial rust should be sought on<br />

plant species infected by microcyclic rusts with<br />

morphologically similar teliospores. This rule,<br />

known as Tranzschel’s Law, has been applied<br />

several times, and DNA-based studies have<br />

confirmed the close relationship, for example<br />

between P. mesnieriana and P. coronata (Zambino &<br />

Szabo, 1993; Shat<strong>to</strong>ck & Preece, 2000). In other<br />

words, the telia of the microcyclic species mimic<br />

the aecia of the macrocyclic ances<strong>to</strong>r. These<br />

two species are then said <strong>to</strong> be correlated. The<br />

specialization of a recently evolved microcyclic<br />

rust on<strong>to</strong> the alternate host of the ances<strong>to</strong>r rust<br />

species may have something <strong>to</strong> do with the<br />

observation that the alternate hosts are almost<br />

always perennial, whereas principal hosts may<br />

be annual (Shat<strong>to</strong>ck & Preece, 2000). Examples of<br />

the reverse case, i.e. the evolution of the microcyclic<br />

species on the principal host of the<br />

macrocyclic ancestral species, do not seem <strong>to</strong> be<br />

known.<br />

22.2.3 The infection process<br />

Concise reviews of this vast <strong>to</strong>pic have been<br />

written by Heath and Skalamera (1997), Mendgen<br />

(1997) and Hahn (2000). One of the differences<br />

between germ tubes arising from basidiospores<br />

and those arising from heterokaryotic spores<br />

(aeciospores or urediniospores) is that the former<br />

usually penetrate the cuticle directly without an<br />

appressorium, whereas the latter usually form<br />

an appressorium and preferentially penetrate<br />

through s<strong>to</strong>mata (Figs. 22.4b, 22.5; Mendgen,<br />

1997). An exception <strong>to</strong> this generalization is<br />

presented by germinating urediniospores of<br />

Phakopsora spp., which show appressoriummediated<br />

direct penetration of the cuticle<br />

(Adendorff & Rijkenberg, 2000). There are also<br />

differences in the carbohydrate polymers making<br />

up the walls of monokaryotic and dikaryotic<br />

stages of the same species of rust fungus, and at<br />

different steps of the infection process (Freytag &<br />

Mendgen, 1991). Relatively little is known about<br />

infection from basidiospores, not least because<br />

these are difficult <strong>to</strong> obtain in the quantity<br />

required for experiments (see Gold & Mendgen,<br />

1991). We shall therefore focus on the infection<br />

process arising from germinating urediniospores,<br />

since these are the most thoroughly<br />

researched system and have a major impact on<br />

agriculture as carriers of the repeated infection<br />

cycle.<br />

Before urediniospores can germinate, germination<br />

au<strong>to</strong>inhibi<strong>to</strong>rs such as methyl-cis-3,<br />

4-dimethoxycinnamate (e.g. in Uromyces appendiculatus)<br />

must be diluted out or degraded. This<br />

substance is active at concentrations in the<br />

10 11 M range (Macko et al., 1970; Staples, 2000),<br />

i.e. at a similarly low concentration as the sex<br />

hormones of Achlya (see p. 86). This makes it one<br />

of the most potent biological molecules. Wolf<br />

(1982) has suggested that the au<strong>to</strong>inhibi<strong>to</strong>r acts<br />

by blocking the lysis of the germ pore. Lysis must


UREDINALES: THE RUST FUNGI<br />

615<br />

Fig 22.3 Attachment of a hydrated urediniospore of<br />

Uromyces viciae-fabae <strong>to</strong> the surface of a broad-bean leaf1h<br />

after contact. (a) Spore removed with sticky tape.The germ<br />

pore has partially lysed. (b) The adhesion pad on the leaf from<br />

which the spore was removed. Some of the germ pore<br />

material has become incorporated in<strong>to</strong> the pad which has<br />

made firm contact with the host cuticle.From Clement et al.<br />

(1997), with permission from Elsevier.Original image kindly<br />

provided by J. A.Clement.<br />

occur before the germ tube can emerge (see<br />

Fig. 22.3).<br />

Attachment of urediniospores <strong>to</strong> surfaces is<br />

a multi-step process. Initial attachment is probably<br />

purely physical and based on hydrophobic<br />

interactions since it is stronger on hydrophobic<br />

than hydrophilic surfaces (Terhune & Hoch,<br />

1993). As soon as the urediniospore becomes<br />

fully hydrated upon contact with water, there is<br />

evidence of the formation of an adhesion pad<br />

of unknown composition (Fig. 22.3; Clement<br />

et al., 1997). Further, cutinases and other esterases<br />

are released, and their activity is thought<br />

<strong>to</strong> modify the surface properties of the host<br />

cuticle, cementing the attachment pad <strong>to</strong> the<br />

host surface (Deising et al., 1992). The germ tube<br />

is also tightly attached <strong>to</strong> the host surface by<br />

means of a glue which probably consists of<br />

glucans and proteins (Epstein et al., 1985;<br />

Chaubal et al., 1991).<br />

The process of appressorium differentiation by<br />

dikaryotic stages of rusts may be unique among<br />

fungi in being triggered by thigmotropism. A<br />

ridge about 0.5 mm high is required and sufficient<br />

<strong>to</strong> induce appressorium differentiation even on<br />

chemically inert surfaces (Fig. 22.4a; Hoch et al.,<br />

1987; Allen et al., 1991). In nature, the relevant<br />

<strong>to</strong>pographic feature is the s<strong>to</strong>matal lip, i.e. the<br />

point where the cuticle broke during the developmental<br />

expansion and opening of the s<strong>to</strong>ma<br />

(Terhune et al., 1991). Firm attachment of the<br />

germ tube <strong>to</strong> the surface is required for the<br />

perception of the signal for appressorium induction<br />

in urediniospore germ tubes of U. appendiculatus,<br />

and there is evidence that the reception<br />

of the physical signal involves microtubules and<br />

integrin-like molecules (Corrêa et al., 1996) and<br />

is transmitted via stretch-activated Ca 2þ channels<br />

(Zhou et al., 1991). We have already come across<br />

stretch-activated Ca 2þ channels and integrin as<br />

possible regula<strong>to</strong>rs of the rate of hyphal tip<br />

extension (see p. 8). Once the signal has been<br />

perceived, an appressorium differentiates in as<br />

little as 60 min.<br />

Following the differentiation of an appressorium<br />

over a s<strong>to</strong>ma, a thin penetration hypha<br />

develops which swells beneath the guard cells<br />

<strong>to</strong> form the subs<strong>to</strong>matal vesicle (Fig. 22.5).<br />

From this, intercellular hyphae grow and form<br />

appressorium-like structures called haus<strong>to</strong>rial<br />

mother cells on the surface of leaf mesophyll<br />

cells. The haus<strong>to</strong>rial mother cell co-ordinates<br />

penetration of the plant cell, leading <strong>to</strong> the<br />

formation of a haus<strong>to</strong>rium. Morphologically<br />

recognizable haus<strong>to</strong>ria are not normally formed<br />

by monokaryotic rust stages; instead, the<br />

hyphae appear <strong>to</strong> grow through mesophyll cells<br />

whose plasmalemma invaginates around them


616 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

Fig 22.4 Appressorium formation by urediniospore germ tubes. (a) The runner bean rust Uromyces appendiculatus.Urediniospores<br />

inoculated on<strong>to</strong> a polystyrene surface with artificial ridges (0.5 mm high) have germinated and differentiated appressoria upon<br />

reaching the ridge. (b) The broad bean rust Uromyces viciae-fabae.The germ tube has followed <strong>to</strong>pographical features of the host<br />

surface until it has reached a s<strong>to</strong>ma, over which an appressorium has formed. (a) Reprinted from Allen et al. (1991) with permission<br />

by APS Press.Original print kindly provided by H.C. Hoch. (b) reprinted from Mendgen (1997), with kind permission of Springer<br />

Science and Business Media.Original print kindly provided by K. Mendgen. Both images approximately <strong>to</strong> same scale.<br />

Fig 22.5 Schematic comparison of infection processes in Uromyces. (a) Germinating basidiospore.The monokaryotic germ tube<br />

infects directly through the epidermal wall without forming a pronounced appressorium.Growth is both inter- and intracellular.<br />

Differentiated haus<strong>to</strong>ria are not formed, and intracellular hyphae are regarded as M-haus<strong>to</strong>ria. (b) Germinating urediniospore.<br />

The dikaryotic germ tube forms an appressorium above a s<strong>to</strong>ma and infects through the s<strong>to</strong>matal opening.Growth in planta is<br />

mainly intercellular. Stalked globose haus<strong>to</strong>ria (D-haus<strong>to</strong>ria) arising from haus<strong>to</strong>rial mother cells are the only intracellular organs<br />

of note. Both M-haus<strong>to</strong>ria and D-haus<strong>to</strong>ria are surrounded by a matrix (not drawn). Based on Mendgen (1997) and Mendgen and<br />

Hahn (2002).<br />

(Gold & Mendgen, 1991). These are sometimes<br />

called M-haus<strong>to</strong>ria. In contrast, differentiated<br />

haus<strong>to</strong>ria are formed by dikaryotic hyphae<br />

(D-haus<strong>to</strong>ria), and valuable physiological work<br />

has been performed on them. This is summarized<br />

in the following section.<br />

22.2.4 Thephysiologyofbiotrophyin<br />

rust fungi<br />

The dikaryotic haus<strong>to</strong>rium of rust fungi<br />

(Fig. 22.7) is functionally very similar <strong>to</strong> that of<br />

the Erysiphales (see Fig. 13.5). The haus<strong>to</strong>rial<br />

mother cell emits a narrow penetration tube


UREDINALES: THE RUST FUNGI<br />

617<br />

Fig 22.6 Cryo SEM of freeze-fractured haus<strong>to</strong>ria<br />

of Uromyces viciae-fabae.(a)Intacthaus<strong>to</strong>rium<br />

(Hau) connected <strong>to</strong> the intercellular haus<strong>to</strong>rial<br />

mother cell (HMC) by a thin penetration tube.The<br />

neck-band (NB) is visible in the penetration tube.<br />

The vacuole (Vac) of the infected plant cell is also<br />

obvious. (b) Fracture through a haus<strong>to</strong>rium,<br />

revealing one of its two nuclei (HN).The nucleus<br />

of the infected plant cell (PN) is closely associated<br />

with the haus<strong>to</strong>rium, a feature frequently<br />

observed in rust infections.The host cell wall<br />

(CW), vacuole (Vac) and a Golgi stack (G) are also<br />

visible. Both images <strong>to</strong> same scale. Previously<br />

unpublished images very kindly provided by<br />

E. Kemen and K. Mendgen.<br />

which swells inside the host cell <strong>to</strong> form the<br />

haus<strong>to</strong>rial body. This contains a full complement<br />

of organelles, including two nuclei. In some<br />

rusts, especially the cereal-infecting species,<br />

these two nuclei sometimes fuse in<strong>to</strong> one diploid<br />

nucleus (Harder & Chong, 1984). From the inside<br />

outwards, the haus<strong>to</strong>rial cy<strong>to</strong>plasm is surrounded<br />

by the haus<strong>to</strong>rial membrane, the<br />

haus<strong>to</strong>rial wall, the extrahaus<strong>to</strong>rial matrix and<br />

the extrahaus<strong>to</strong>rial membrane (i.e. the modified<br />

plant plasmalemma). The haus<strong>to</strong>rial matrix is<br />

sealed against the apoplast outside the infected<br />

plant cell by means of a neckband (Fig. 22.6a).<br />

The host nucleus is often closely associated with<br />

the extrahaus<strong>to</strong>rial membrane (Fig. 22.6b).<br />

As in the Erysiphales (see p. 398), the extrahaus<strong>to</strong>rial<br />

membrane surrounding D-haus<strong>to</strong>ria<br />

seems <strong>to</strong> lack ATPase activity (Baka et al., 1995),<br />

thus indicating that the infected plant cell has no<br />

effective means <strong>to</strong> restrict the efflux of metabolites<br />

in<strong>to</strong> the haus<strong>to</strong>rial matrix. In contrast,<br />

ATPase activity in the fungal haus<strong>to</strong>rial<br />

membrane may actually be increased relative <strong>to</strong><br />

normal hyphae (Struck et al., 1996). Not surprisingly,<br />

pro<strong>to</strong>n-driven hexose and amino acid<br />

uptake appears <strong>to</strong> occur from the matrix across<br />

the haus<strong>to</strong>rial membrane in<strong>to</strong> the haus<strong>to</strong>rium<br />

(Fig. 22.7; Voegele & Mendgen, 2003). The uptake<br />

mechanism is thus equivalent <strong>to</strong> the uptake of<br />

solutes in<strong>to</strong> growing hyphae (see Fig. 1.11).<br />

Whereas powdery mildews appear <strong>to</strong> rely on the<br />

host plant for the hydrolysis of the transport<br />

disaccharide sucrose in<strong>to</strong> the hexoses fruc<strong>to</strong>se<br />

and glucose prior <strong>to</strong> uptake in<strong>to</strong> the haus<strong>to</strong>rium<br />

(p. 398), Voegele and Mendgen (2003) have<br />

suggested that the rust haus<strong>to</strong>rium secretes


618 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

Fig 22.7 Diagram showing nutrient uptake mechanisms in a D-haus<strong>to</strong>rium. Enzymatic reactions and pro<strong>to</strong>n-driven pumping<br />

are indicated by solid arrows; dotted arrows indicate diffusion or translocation processes. Sucrose (Sucr.) is hydrolysed <strong>to</strong> fruc<strong>to</strong>se<br />

(Fruct.) and glucose (Gluc.) by invertase (Inv.) secreted in<strong>to</strong> the extrahaus<strong>to</strong>rial matrix.These monosaccharides as well as amino<br />

acids (AA) are taken up across the haus<strong>to</strong>rial membrane by specialized porter proteins fuelled by a transmembrane H þ gradient.<br />

Fruc<strong>to</strong>se is converted <strong>to</strong> the fungal transport compound manni<strong>to</strong>l (Mann.) by alcohol dehydrogenase within the haus<strong>to</strong>rium.<br />

Modified and redrawn fromVoegele and Mendgen (2003).<br />

invertase in<strong>to</strong> the matrix <strong>to</strong> perform the necessary<br />

hydrolysis (Fig. 22.7).<br />

All rusts are obligately biotrophic in nature,<br />

i.e. they need <strong>to</strong> parasitize living host plants in<br />

order <strong>to</strong> complete their life cycle. The fact that<br />

many of them can now be cultivated on agar<br />

media in the labora<strong>to</strong>ry does not alter their<br />

status as obligate biotrophs in nature (P. G.<br />

Williams, 1984). The initial report of the cultivation<br />

of P. graminis on a simple medium (Williams<br />

et al., 1966) raised high hopes for a breakthrough<br />

in research on rusts, but interest has waned in<br />

recent years. Excellent accounts recalling the<br />

excitement of discovery have been written by<br />

Maclean (1982) and P. G. Williams (1984). It is<br />

now possible <strong>to</strong> cultivate rust fungi on relatively<br />

simple agar media based on minerals (e.g.<br />

Czapek-Dox agar) with sucrose or glucose as<br />

carbon source and yeast extract, pep<strong>to</strong>ne or<br />

certain amino acids as a nitrogen source.<br />

22.2.5 Host resistance<br />

Resistance against rust infections can manifest<br />

itself at different stages of the infection process.<br />

Most commonly it becomes evident as a hypersensitive<br />

response when the infectious hypha<br />

attempts <strong>to</strong> breach the first host cell wall (Heath<br />

& Skalamera, 1997). With monokaryotic stages of<br />

rust fungi, this is the direct penetration through<br />

the wall of the epidermal cell, whereas the<br />

resistance response <strong>to</strong> an incompatible dikaryotic<br />

strain occurs later, during or after the


UREDINALES: THE RUST FUNGI<br />

619<br />

formation of the first haus<strong>to</strong>rium (Heath, 1982,<br />

2002). Either way, the hypersensitive response<br />

triggers biochemical events involved in local and<br />

systemic defence against the pathogen (Heath,<br />

2000). One point of difference between penetration<br />

by rust fungi and powdery mildews is that<br />

the host plant mounts some resistance response<br />

against the latter even if the interaction ultimately<br />

turns out <strong>to</strong> be compatible, whereas no<br />

such recognition occurs if a compatible rust<br />

germ tube penetrates (Heath, 2002). The ultrastructural<br />

consequence is that powdery mildew<br />

haus<strong>to</strong>ria commonly have a callose ring around<br />

their neck, whereas this is absent or reduced in<br />

the case of rust haus<strong>to</strong>ria (Heath & Skalamera,<br />

1997).<br />

The gene-for-gene concept was proposed <strong>to</strong><br />

explain the specific interactions between the<br />

rust Melampsora lini and its host, Linum usitatissimum<br />

(flax). It postulates that for every resistance<br />

gene of the host plant there is a matching<br />

virulence gene in the pathogen (see Flor, 1971).<br />

Resistance is usually dominant (R) whereas<br />

virulence is recessive (a). An incompatible interaction<br />

results if a rust fungus with an avirulence<br />

allele (A) attempts <strong>to</strong> infect a host carrying the<br />

matching resistance (R) allele. The identity of<br />

most molecules interacting in recognition is<br />

still unknown. However, Catanzariti et al. (2006)<br />

have demonstrated that the protein products of<br />

several avirulence genes are secreted by developing<br />

haus<strong>to</strong>ria of M. lini, and that these seem <strong>to</strong><br />

enter the cy<strong>to</strong>plasm of infected host cells where<br />

recognition leading <strong>to</strong> hypersensitivity occurs.<br />

The localization of proteins of haus<strong>to</strong>rial origin<br />

in host cells, including the host nucleus, was also<br />

demonstrated by Kemen et al. (2005) for Uromyces<br />

spp. on broad bean (Vicia faba). These authors<br />

suggested that fungal avirulence genes might<br />

encode transcription fac<strong>to</strong>rs involved in manipulating<br />

the host’s metabolism during the<br />

biotrophic interaction. In this theory, avirulence<br />

would result if the host cell managed <strong>to</strong> detect<br />

the fungal transcription fac<strong>to</strong>r as foreign. It is<br />

not yet known how the avirulence proteins are<br />

translocated in<strong>to</strong> the plant cell. The molecular<br />

basis of gene-for-gene interactions involving<br />

rust fungi is therefore fundamentally<br />

different from that, for example, in<br />

Cladosporium fulvum infecting <strong>to</strong>ma<strong>to</strong>, where<br />

recognition events occur at the host plasma<br />

membrane (see p. 482).<br />

Rusts, like most other fungi, possess an<br />

uncanny ability <strong>to</strong> overcome major gene resistance<br />

based on gene-for-gene interactions, especially<br />

if a single resistance gene is involved.<br />

Genetic variation is enhanced by the ability <strong>to</strong><br />

reproduce sexually in the field if the alternate<br />

host is available. Even in the absence of the<br />

alternate host, however, rust fungi can still<br />

undergo genetic recombination by anas<strong>to</strong>mosis<br />

and nuclear exchange in various other ways (see<br />

p. 625). Different races of a given pathogen can<br />

be distinguished by their ability <strong>to</strong> infect any<br />

host in a set of cultivars containing defined<br />

resistance genes alone or in combination. Such<br />

tests are routinely employed by plant pathologists<br />

for the identification of races, and for<br />

moni<strong>to</strong>ring their spread (see p. 626).<br />

Major gene resistance against rust fungi was<br />

recognized early in the twentieth century by<br />

Biffen (1905) and others, and extensive breeding<br />

programmes were initiated. Typically a new<br />

cultivar produces excellent results for a few<br />

cropping seasons until resistant races develop<br />

and spread in the field, thereby rendering the<br />

breeders’ efforts futile. This is the ‘boom-andbust’<br />

cycle. The ‘bust’ of a cultivar can be delayed<br />

if it contains a ‘pyramid’ of several resistance<br />

genes which the pathogen may be unable <strong>to</strong><br />

overcome. Sometimes the breeding for resistance<br />

against one pathogen can lead <strong>to</strong> susceptibility<br />

<strong>to</strong> another, as in the case of the Vic<strong>to</strong>ria oat<br />

cultivar which showed good resistance against<br />

P. coronata f. sp. avenae but was devastated by<br />

Cochliobolus vic<strong>to</strong>riae (see p. 471). In addition<br />

<strong>to</strong> major gene resistance which is commonly<br />

associated with the hypersensitive response,<br />

there are other types of resistance. Some plant<br />

genes do not afford <strong>to</strong>tal resistance but give<br />

a moderate degree of resistance. If a combination<br />

of several genes is involved, the resistance<br />

may well be more durable in the field than<br />

single-gene resistance. An example of such ‘field’<br />

or ‘partial’ resistance is the reduced formation<br />

of appressoria over s<strong>to</strong>mata of grasses covered<br />

by a particularly thick wax layer, which seems<br />

<strong>to</strong> mask the <strong>to</strong>pographic features of the


620 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

s<strong>to</strong>matal surface acting as the thigmotropic<br />

signal (Rubiales & Niks, 1996). There is also<br />

adult-plant resistance, usually an interaction<br />

between various genes which retard the progress<br />

of infection <strong>to</strong> such a degree that the losses are<br />

reduced <strong>to</strong> economically acceptable levels.<br />

22.3 Puccinia graminis, the cause of<br />

black stem rust<br />

A connection between the barberry bush (Berberis<br />

vulgaris) and rust disease on wheat and other<br />

cereals had been suspected for centuries by<br />

farmers who noticed the frequent occurrence of<br />

crop damage around or downwind from barberry<br />

bushes. Further, an al<strong>to</strong>gether different-looking<br />

fungus known as Aecidium berberidis was known<br />

<strong>to</strong> infect barberry leaves. The formal proof that<br />

the barberry and wheat pathogens were different<br />

stages of the same species was made by An<strong>to</strong>n de<br />

Bary who, in 1865, performed cross-inoculation<br />

experiments <strong>to</strong> show that basidiospores derived<br />

from germinating teliospores from wheat were<br />

able <strong>to</strong> cause spermogonial and aecial infections<br />

on the barberry. This s<strong>to</strong>ry has been recounted<br />

many times, but nobody has <strong>to</strong>ld it better than<br />

Large (1940).<br />

Black stem rust affects especially wheat but<br />

also most other cereals and a range of wild<br />

grasses. Crop losses can be severe as a reduction<br />

in quantity as well as quality of the grain yield.<br />

The host epidermis can become ruptured over<br />

much of its surface in severe infections, thereby<br />

debilitating the plants (see Fig. 22.8d). Serious<br />

epidemics have occurred in the past, e.g. in the<br />

USA in 1904 and especially during the war year<br />

1916. In fact, losses were so severe (up <strong>to</strong> 50% in<br />

the Great Plains; Eversmeyer & Kramer, 2000)<br />

that P. graminis was seriously considered and<br />

developed as a biological weapon by the US<br />

Government during the 1960s (Line & Griffith,<br />

2001). On susceptible wheat cultivars and without<br />

chemical protection, P. graminis can cause<br />

<strong>to</strong>tal crop failure.<br />

Although it is an ecologically obligate<br />

biotroph, Puccinia graminis is able <strong>to</strong> infect an<br />

as<strong>to</strong>nishingly wide range of grass and cereal<br />

hosts. Gäumann (1959) listed 365 host species in<br />

Fig 22.8 Puccinia graminis. (a) Spermogonial pustules on the upper surface of a leaf of Berberis vulgaris.Notethedropsofnectar.<br />

(b) Aecia on the underside of a Berberis leaf.The outer frilly layer is the white peridium, within which is a mass of orange-coloured<br />

aeciospores. (c) Wheat leaf showing uredinia which appear as reddish-brown powdery masses. (d) Wheat straw showing telia as<br />

black raised pustules.


PUCCINIA GRAMINIS, THE CAUSE OF BLACK STEM RUST<br />

621<br />

54 genera, and this list is still expanding.<br />

However, no single isolate of P. graminis is able<br />

<strong>to</strong> infect all these host species. Instead, the<br />

species P. graminis can be separated in<strong>to</strong> several<br />

specialized forms which have become adapted <strong>to</strong><br />

one or a few principal host species. These are<br />

termed formae speciales (sing. forma specialis,<br />

abbreviated as ‘f. sp.’) because they cannot be<br />

distinguished reliably from each other on morphological<br />

criteria. A morphologically distinct<br />

form of the same species which can be clearly<br />

identified by microscopy or other means would<br />

be called varietas (abbreviated as ‘var.’).<br />

The most important formae speciales are<br />

P. graminis f. sp. tritici (on wheat), f. sp. avenae<br />

(on oat) and f. sp. secalis (on rye). In addition, there<br />

are several forms on wild grasses, e.g. f. sp. phleipratensis,<br />

f. sp. lolii and f. sp. agrostidis (Wilson &<br />

Henderson, 1966; Anikster, 1984). It is possible <strong>to</strong><br />

produce hybrids between some of these formae<br />

speciales, e.g. between f. sp. tritici and f. sp. secalis<br />

(Green, 1971). The hybrid aeciospores are not very<br />

virulent on either principal host, and Green<br />

(1971) has argued that the hybrids resemble a<br />

more primitive form of Puccinia graminis with low<br />

virulence and a wide host range, and that evolution<br />

in stem rust of cereals is progressing from<br />

low virulence and a wide host range <strong>to</strong> high<br />

virulence and a narrowed host range.<br />

Each forma specialis on cereals in turn forms<br />

hundreds of races distinguished by the infection<br />

responses of differential host cultivars, with<br />

new races continually evolving (see p. 626). This<br />

feature highlights the remarkable genetic and<br />

physiological flexibility of rust fungi.<br />

22.3.1 Puccinia graminis on barberry<br />

The basidiospores are released in spring from<br />

overwintered cereal stubble at about the time<br />

when fresh barberry leaves unfold. Basidiospores<br />

are able <strong>to</strong> germinate by repetition if they do not<br />

land on a suitable host surface (Fig. 22.10e).<br />

Infection gives rise <strong>to</strong> a haploid monokaryotic<br />

mycelium which shows inter- and intra-cellular<br />

growth and colonizes the host tissue extensively.<br />

Generally, monokaryotic stages of rust fungi<br />

show more widespread colonization of host<br />

tissue than their dikaryotic counterparts.<br />

Viewed from the surface, the colonized barberry<br />

leaf area appears as a yellowish circular lesion.<br />

On the upper surface of this lesion, several<br />

flask-shaped spermogonia develop whose necks<br />

protrude beyond the epidermal layer. Among<br />

the orange-coloured tapering periphyses surrounding<br />

the opening of each spermogonium<br />

are several thinner, hyaline branched hyphae,<br />

the flexuous (or receptive) hyphae. Lining the<br />

inside surface of the spermogonium are tapering<br />

annellides which give rise <strong>to</strong> small uninucleate<br />

spermatia. These ooze out through the mouth<br />

of the spermogonium and are held by the periphyses<br />

in a drop of sticky sweet-smelling liquid<br />

(Figs. 22.8a, 22.9a). Within the mesophyll of the<br />

barberry leaf, the haploid mycelium gives rise<br />

<strong>to</strong> several spherical structures called pro<strong>to</strong>-aecia.<br />

These are mostly made up of large-celled pseudoparenchyma,<br />

but in the upper region is a cap<br />

of smaller, denser cells (Fig. 22.9b).<br />

Single haploid lesions are incapable of<br />

further development unless cross-fertilization<br />

occurs. The sweet-smelling spermatial exudate<br />

contains fruc<strong>to</strong>se and several volatile substances<br />

(see p. 629) which attract insects feeding on the<br />

nectar and carrying the spermatia around by<br />

visiting several distinct pustules. The haploid<br />

pustules are of either of the two mating types,<br />

(þ) or( ), and if a (þ) spermatium is brought<br />

close <strong>to</strong> a flexuous hypha of opposite mating<br />

type, it produces a short germ tube which<br />

anas<strong>to</strong>moses with the flexuous hypha (Craigie,<br />

1927; Buller, 1950). Nuclear transfer is followed<br />

by repeated division and migration of the<br />

introduced nucleus <strong>to</strong>wards the pro<strong>to</strong>-aecium<br />

(Craigie & Green, 1962). This results in the<br />

dikaryotization of the haploid mycelium until<br />

binucleate cells become visible in the cap region<br />

of the pro<strong>to</strong>-aecium after about 3 days. The<br />

binucleate cells now start <strong>to</strong> give rise <strong>to</strong> chains<br />

of alternating long and short cells which are also<br />

binucleate. The longer cells enlarge and become<br />

aeciospores, but the shorter cells disintegrate as<br />

the spore chains develop (Fig. 22.9c). During the<br />

development of the spore chains, the large<br />

pseudoparenchyma<strong>to</strong>us cells of the pro<strong>to</strong>aecium<br />

are also crushed and pushed aside.<br />

Surrounding the chains of spores is a specially<br />

differentiated layer of cells homologous with the


622 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

Fig 22.9 Puccinia graminis. (a) T.S. spermogonium on a leaf of Berberis vulgaris.The spermogonium has penetrated the upper<br />

epidermis.The wall of the spermogonium is lined by tapering annellides which give rise <strong>to</strong> spermatia. (b) T.S. leaf of B. vulgaris<br />

showing a spermogonium and a pro<strong>to</strong>-aecium. (c) T.S. leaf of B. vulgaris showing an aecium in section.The aecium has burst through<br />

the lower epidermis of the host leaf. Note the columns consisting of alternating large and small cells.The large cells are the<br />

aeciospores.<br />

spore chains, whose outer walls are thick and<br />

fibrous. This layer forms a clearly defined border<br />

or peridium surrounding the spores. Eventually<br />

the peridium and spore chains burst through the<br />

lower epidermis. The peridium ruptures and<br />

aeciospores are now visible as orange-coloured<br />

cells enclosed by the white cup-like peridium<br />

(Fig. 22.8b). Since the aecia are commonly seen<br />

clustered <strong>to</strong>gether beneath a spermogonial<br />

lesion, this stage is popularly known as the<br />

cluster-cup stage. In a section through the centre<br />

of a group of young aecia, it is usually possible <strong>to</strong><br />

find the spermogonia penetrating the upper<br />

epidermis and the aecia penetrating the lower.<br />

The aeciospores are violently projected from the<br />

end of the spore chain by rounding off of the<br />

flattened interface between adjacent spores<br />

(Ingold, 1971). Aeciospores are unable <strong>to</strong> reinfect<br />

barberry but readily infect wheat or<br />

other grasses.<br />

It is relatively easy <strong>to</strong> establish and maintain<br />

P. graminis in garden situations by planting<br />

B. vulgaris and allowing Agropyron repens <strong>to</strong> grow<br />

underneath (Webster et al., 1999). The form most<br />

likely <strong>to</strong> grow is P. graminis f. sp. secalis which<br />

readily alternates between Berberis and grasses<br />

in Britain (Wilson & Henderson, 1966). In other<br />

countries the connection between Berberis and<br />

P. graminis f. sp. tritici is strong, and the first<br />

barberry eradication laws were implemented<br />

long before de Bary had formally proven the<br />

connection. The first recorded cases of such laws<br />

were in Rouen in 1660 and Massachusetts in 1755.<br />

The greatest effort by far was the barberry<br />

eradication campaign undertaken as a consequence<br />

of the 1916 epidemic in the United States.<br />

Widespread eradication of susceptible Berberis<br />

spp. started in 1918 and continued on a massive<br />

scale well in<strong>to</strong> the 1930s, and locally for several<br />

decades afterwards. Overall, this campaign is<br />

considered <strong>to</strong> have been successful, with wheat<br />

rust epidemics much reduced in severity, especially<br />

on a local scale. Because aeciospores are not<br />

particularly long-lived and are not produced in<br />

vast numbers, the direct impact of barberry<br />

bushes on wheat crops is limited <strong>to</strong> less than<br />

2 miles. Perhaps more importantly, barberry<br />

eradication also retarded the evolution of new<br />

races of P. graminis due <strong>to</strong> the reduced ability of<br />

the pathogen <strong>to</strong> reproduce sexually. Campbell<br />

and Long (2001) have provided a highly readable<br />

account of the eradication campaign in the USA.


PUCCINIA GRAMINIS, THE CAUSE OF BLACK STEM RUST<br />

623<br />

22.3.2 Puccinia graminis on cereals<br />

A symp<strong>to</strong>m of infection on wheat leaves and<br />

stems is the appearance of brick-red pustules<br />

(uredinia) between the veins. Uredinia contain<br />

stalked, one-celled dikaryotic urediniospores<br />

which burst through the epidermis (Figs. 22.8c<br />

and 22.10a,b). These have a spiny wall which has<br />

four thinner areas (germ pores) near the middle<br />

of the spore. The urediniospores are detached by<br />

wind and blown <strong>to</strong> fresh wheat leaves upon<br />

which they germinate by extruding a germ tube<br />

from one of the germ pores. Germination<br />

requires free water (e.g. night-time dew) and<br />

proceeds optimally at about 20°C. Infection<br />

Fig 22.10 Puccinia graminis f. sp. secalis. (a) T.S. stem through a uredinium.The stalked unicellular urediniospores are protruding<br />

through the ruptured host epidermis. A teliospore (t) has also been formed. (b) Higher-power detail of urediniospores. Note the<br />

germ pores (g) and the haus<strong>to</strong>ria (h) in the host cells. (c) Germination of urediniospores on host leaf. Note the directional growth<br />

of the germ tubes perpendicular <strong>to</strong> the long axes of the epidermal cells, <strong>to</strong>wards the s<strong>to</strong>mata. (d) T.S. leaf sheath through a telium.<br />

The stalked teliospores are projecting through the ruptured epidermis.Drawing <strong>to</strong> same scale as (a). (e) Germination of teliospores<br />

<strong>to</strong> form metabasidia bearing sterigmata and basidiospores.One basidiospore is giving rise <strong>to</strong> a secondary spore.


624 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

follows the usual pattern for dikaryotic propagules,<br />

i.e. via an appressorium through a s<strong>to</strong>ma<br />

as shown in Fig. 22.5b. Within about 7 21 days<br />

of infection, a new crop of urediniospores is<br />

formed so that inoculum can build up rapidly<br />

within a crop. A single uredinium may produce<br />

between 50 000 <strong>to</strong> 400 000 spores, and there may<br />

be 4 5 generations of urediniospores in the<br />

growing period of a wheat crop. Disease development<br />

occurs best at daytime temperatures<br />

around 30°C. This may explain why stem rust<br />

of wheat is particularly prevalent in areas with<br />

a continental climate, whereas it is not a serious<br />

disease, for example, in Britain.<br />

Puccinia graminis f. sp. tritici can survive the<br />

winter in the mild climates, e.g. of Mexico and the<br />

Southern USA (see Fig. 22.11) as uredinial infections<br />

on its principal host, i.e. wheat crops and<br />

volunteer plants. Despite the almost complete<br />

elimination of the barberry, infections with<br />

P. graminis f. sp. tritici still occur every year in<br />

North America from uredinial lesions overwintering<br />

in the South. Urediniospores are longer-lived<br />

than aeciospores, and waves of inoculum can<br />

move northwards with the prevailing southerly<br />

winds. The track taken is called the ‘Puccinia<br />

pathway’ (Agrios, 2005). Using small aeroplanes,<br />

urediniospores have been detected at altitudes<br />

as high as 5000 feet. The <strong>to</strong>tal distance travelled<br />

each year from the Southern USA <strong>to</strong> Canada<br />

can be in excess of 2000 miles (Stakman &<br />

Christensen, 1946), but in most cases migration<br />

occurs in several shorter intervals, with new crops<br />

of uredinia being produced en route (Eversmeyer &<br />

Kramer, 2000). Consequently, infections start<br />

at successively later dates the further north<br />

the epidemic moves (Fig. 22.11). Attempts have<br />

been made <strong>to</strong> interrupt this migration by planting<br />

cultivars with different resistance genes in<br />

different regions of the Puccinia pathway (Frey<br />

et al., 1973). Towards the end of the season there<br />

may be a reversal of the flow of air, so that wheat<br />

Fig 22.11 The‘Puccinia pathway’ in North America, with dates indicating the average annual arrival of P. graminis f. sp. tritici in<br />

the wheat fields.The fungus regularly overwinters in the uredinial state south of 30° Northern latitude, and occasionally as far north<br />

as 34°. Redrawn from Roelfs (1986) and several other sources.


PUCCINIA GRAMINIS, THE CAUSE OF BLACK STEM RUST<br />

625<br />

in Texas and Mexico may be infected from<br />

urediniospores derived from the northern areas<br />

(Pady & Johns<strong>to</strong>n, 1955). Similar long-distance<br />

transport of urediniospores has been reported<br />

from Europe and India (Nagarajan & Singh, 1990).<br />

Late in the season, uredinial lesions gradually<br />

turn <strong>to</strong> producing teliospores instead of urediniospores.<br />

Telial pustules appear as black raised<br />

streaks along leaf-sheaths and stems of infected<br />

plants (Fig. 22.8d). Teliospores are initially<br />

binucleate, but soon the two nuclei fuse and<br />

the spores survive the winter in the diploid<br />

state. A period of maturation corresponding <strong>to</strong><br />

winter dormancy is required before teliospores<br />

are competent <strong>to</strong> germinate. When they do<br />

germinate, each of the two cells emits a fourcelled<br />

promycelium or metabasidium. Meiosis<br />

gives rise <strong>to</strong> four nuclei which migrate in<strong>to</strong><br />

the four basidiospores and then perform one<br />

further mi<strong>to</strong>tic division. Each cell of the metabasidium<br />

bears one basidiospore containing two<br />

nuclei of the same mating type. The basidiospores<br />

are short-lived and do not travel far, and the<br />

teliospore state is of no significance in an<br />

agricultural context if no barberry bushes are in<br />

the vicinity.<br />

22.3.3 Resistance breeding and<br />

physiological races of P. graminis<br />

f. sp. tritici<br />

If the spores of a given isolate of P. graminis<br />

f. sp. tritici are inoculated on<strong>to</strong> a range of wheat<br />

cultivars, these hosts will differ in their response.<br />

Some may prove <strong>to</strong> be resistant, others highly<br />

susceptible, whilst yet others may be intermediate<br />

in their reaction. Spores from a second<br />

source may give an entirely different pattern<br />

of response. Using the reactions of different<br />

wheat cultivars, it has proven possible <strong>to</strong> classify<br />

P. graminis f. sp. tritici in<strong>to</strong> over 300 physiological<br />

races. The existence of such a large number of<br />

races may be due <strong>to</strong> several fac<strong>to</strong>rs.<br />

1. Sexual recombination. New races of rust<br />

are often found adjacent <strong>to</strong> barberry bushes.<br />

Further, following inoculation with a single race,<br />

a number of variants may be found among the<br />

aeciospore progeny.<br />

2. Anas<strong>to</strong>mosis on the principal host. If two<br />

uredinial mycelia growing on the same host leaf<br />

come in<strong>to</strong> hyphal contact, anas<strong>to</strong>mosis may<br />

result. This can lead <strong>to</strong> the exchange of genetic<br />

material known as somatic hybridization. The<br />

simplest way is the exchange of entire nuclei<br />

which will lead <strong>to</strong> the formation of new heterokaryons<br />

and their dissemination as urediniospores<br />

(Ellingboe, 1961). However, when two<br />

different races of P. graminis were inoculated<br />

simultaneously on<strong>to</strong> a susceptible wheat cultivar,<br />

at least 15 different races were identified<br />

among the urediniospore progeny, which is more<br />

than would be expected from simple nuclear rearrangement<br />

(Watson & Luig, 1958; Bridgmon,<br />

1959). Puccinia graminis contains a haploid set of<br />

18 chromosomes (Boehm et al., 1992), and the<br />

exchange of any of them in the dikaryotic<br />

mycelium during synchronous nuclear division<br />

is conceivable (Hartley & Williams, 1971).<br />

Additionally, parasexual recombination between<br />

chromosomes could occur. Park et al. (1999) have<br />

provided evidence of the occurrence of somatic<br />

hybridization in P. triticina in the field.<br />

3. New races may arise by mutation, as shown<br />

by Park et al. (1995) for the brown leaf rust of<br />

wheat, P. triticina.<br />

The existence of this large range of rust races<br />

complicates the task of the plant breeder, but<br />

fortunately all these races are not prevalent in an<br />

area at any one time. Therefore the breeding of<br />

resistant cultivars remains practicable and profitable<br />

(Johnson, 1953; Dyck & Kerber, 1985). The<br />

frequency of different rust races varies over the<br />

years, reflecting the changes in the wheat<br />

cultivars planted (Fig. 22.12). A new race may<br />

suddenly build up in frequency, as shown by the<br />

appearance of Race 15 (actually a sub-race or<br />

biotype referred <strong>to</strong> as Race 15B) in Canada. The<br />

appearance of new races presents a problem <strong>to</strong><br />

plant breeders, and it is clear that the breeding<br />

and release of resistant cultivars has <strong>to</strong> be<br />

carefully co-ordinated.<br />

The original scheme for using host differentials<br />

was devised by Stakman and Levine (1922)<br />

who used 12 different wheat cultivars (‘standard<br />

differentials’) <strong>to</strong> discriminate between the physiological<br />

races (‘standard races’). The standard<br />

differentials frequently failed <strong>to</strong> reveal important


626 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

Fig 22.12 Diagrammatic representation of the distribution in<br />

Canada of eight physiological races of wheat stem rust during<br />

the period 1919 1951 (after Johnson,1953).<br />

changes in the virulence of the rust population,<br />

and this led <strong>to</strong> the use of supplementary<br />

differentials <strong>to</strong> differentiate between further<br />

physiological races. Gradually the nomenclature<br />

for numbering rust races became quite<br />

complicated. In the wake of the acceptance of<br />

the gene-for-gene concept (see Flor, 1971), a<br />

more rational way of identifying the<br />

physiological races was <strong>to</strong> base their classification<br />

on the resistance genes of the host plant. Using<br />

the wheat cultivar Marquis, several ‘single gene<br />

lines’, i.e. wheat cultivars carrying single resistance<br />

genes, have been generated. The resistance<br />

genes were given the symbols Sr1, Sr2, etc.<br />

(Sr for ‘stem rust’). Some 40 resistance genes<br />

are known (Roelfs, 1985). Efforts at characterizing<br />

pathogen races are now standardized by<br />

means of an international nomenclatural system<br />

(Roelfs & Martens, 1988). An authoritative<br />

account of the his<strong>to</strong>ry and importance of rust


OTHER CEREAL RUSTS<br />

627<br />

race characterization has been written by Roelfs<br />

(1984).<br />

The serious race 15B epidemics of the early<br />

1950s (see Fig. 22.12) resulted in the replacement<br />

of the then prevalent cultivars by others such as<br />

‘Selkirk’, which carried several resistance genes<br />

and remained resistant <strong>to</strong> P. graminis f. sp. tritici<br />

in North America for several decades (Roelfs,<br />

1985). Several wheat varieties more recently<br />

released have also retained their resistance for<br />

many years in the field (Roelfs, 1984; Dyck &<br />

Kerber, 1985), so that no major stem rust<br />

epidemic has occurred in the USA during the<br />

past 50 years. This success is probably due <strong>to</strong> two<br />

main reasons, namely the combination of several<br />

resistance genes in one cultivar which the rust<br />

fungus is apparently unable <strong>to</strong> overcome, and<br />

the stabilization of rust races by the elimination<br />

of sexual reproduction as a consequence of the<br />

barberry eradication campaign.<br />

22.4 Other cereal rusts<br />

Although potentially the most damaging cereal<br />

rust, P. graminis has lost much of its menace<br />

for reasons discussed in the preceding section.<br />

Several other rust species now cause more<br />

serious crop losses than P. graminis. These are<br />

briefly discussed below. Cereal rusts are primarily<br />

controlled by breeding of resistant cereal<br />

cultivars, and this is an ongoing battle against<br />

new races which, once arisen, are capable of<br />

spreading rapidly and on a global scale (Chen,<br />

2005; Kolmer, 2005). Chemical control by the<br />

application of fungicides is practised <strong>to</strong> protect<br />

crops grown for seed production, and <strong>to</strong> control<br />

severe outbreaks of rust if resistance fails. <strong>Fungi</strong>cides<br />

used include the systemic strobilurin-type<br />

compounds and the ergosterol biosynthesisinhibiting<br />

triazoles (see Fig. 13.15), as well as<br />

various protectant compounds.<br />

22.4.1 Puccinia triticina (brown leaf<br />

rust of wheat)<br />

Puccinia triticina was formerly named P. recondita<br />

f. sp. tritici, but Zambino and Szabo (1993) and<br />

Anikster et al. (1997) have shown that it is not<br />

closely related <strong>to</strong> P. recondita, which infects rye<br />

but is only of minor significance. Puccinia triticina<br />

has displaced P. graminis as the economically<br />

most important rust of wheat especially in North<br />

America and Eastern Europe (Samborski, 1985).<br />

This species is macrocyclic, with Thalictrum speciosissimum<br />

(¼ T. flavum; Ranunculaceae) serving as<br />

the alternate host, although it seems <strong>to</strong> survive<br />

mainly in the uredinial state. Its epidemiology is<br />

therefore similar <strong>to</strong> that of P. graminis f. sp. tritici<br />

(Eversmeyer & Kramer, 2000). The temperature<br />

optima for urediniospore germination and infection<br />

(about 18°C) and sporulation (25°C) are<br />

lower than those of P. graminis (Singh et al.,<br />

2002). Teliospores are not readily formed, and<br />

the disease symp<strong>to</strong>ms are easily recognized by<br />

the chocolate-brown uredinial lesions on wheat<br />

leaves. There is no infection of the stem, in contrast<br />

<strong>to</strong> P. graminis. Triticale, a hybrid between<br />

wheat (Triticum) and rye (Secale), is also affected.<br />

About 50 Lr (‘leaf rust’) resistance genes are<br />

known, and wheat leaf rust is currently controlled<br />

mainly by using cultivars containing<br />

a pyramid of several of these Lr genes.<br />

22.4.2 Puccinia striiformis (stripe rust or<br />

yellow rust on wheat and barley)<br />

The alternate host of this rust species has not<br />

yet been found, and it may no longer have one.<br />

Therefore, its life cycle is probably confined <strong>to</strong><br />

the grass or cereal hosts on which uredinia and<br />

telia are produced. This rust can be distinguished<br />

from P. triticina by its yellow uredinia which are<br />

arranged in stripe-like rows following the leaf<br />

veins. There are several formae speciales, but their<br />

host ranges overlap <strong>to</strong> a greater extent than<br />

in the other cereal rusts and they are therefore<br />

less clearly delimited. The most important<br />

forms are those on wheat (f. sp. tritici) and<br />

barley (f. sp. hordei). The temperature optima of<br />

P. striiformis are the lowest of any of the common<br />

cereal rusts, with urediniospore germination<br />

and penetration around 10°C, and urediniospore<br />

production at 12 15°C (Singh et al., 2002). This<br />

species overwinters by using winter crops and<br />

volunteers as a ‘green bridge’, and can cause<br />

early epidemics in spring. Zadoks and Bouwman<br />

(1985) have estimated that a single uredinium


628 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

per hectare is sufficient <strong>to</strong> spark an epidemic.<br />

Even if no lesions are seen on winter wheat, the<br />

rust may still be present, surviving as latent<br />

infections. Crop losses can be severe in the cooler<br />

climates of North-Western Europe and North<br />

America where winter cereals, especially winter<br />

wheat, are widely grown (Stubbs, 1985). Like<br />

most other rusts, P. striiformis is controlled<br />

mainly by resistance breeding (Line, 2002). Chen<br />

(2005) has given a detailed account of recent<br />

epidemics and migrations of P. striiformis.<br />

22.4.3 Puccinia hordei (barley leaf rust or<br />

‘dwarf rust’)<br />

This macrocyclic rust alternates between<br />

Ornithogalum spp. (Hyacinthaceae) as the alternate<br />

host and cultivated and wild barley (Hordeum)<br />

species as principal host. Clifford (1985) considered<br />

this species <strong>to</strong> be the most important rust on<br />

barley. The uredinial lesions are brownish in<br />

colour and are smaller than those of other rusts,<br />

hence the name ‘dwarf rust’. In agricultural<br />

situations, the alternate host is probably of little<br />

importance as the fungus can overwinter in the<br />

uredinial state on volunteer plants or winter<br />

barley. Its optimum temperatures for infection<br />

(15°C) and sporulation (10 20°C) are almost as<br />

low as those for P. striiformis. In Europe in the<br />

1970s, P. hordei caused major crop losses because<br />

the cultivars sown were highly susceptible.<br />

However, the situation seems <strong>to</strong> have improved<br />

in recent years, and several cultivars showing<br />

good partial resistance have been bred (Niks et al.,<br />

2000).<br />

22.4.4 Puccinia coronata (crown rust of oats<br />

and forage grasses)<br />

This rust is so named because of the spiny<br />

extensions at the apex of its teliospore (see<br />

Fig. 22.13a). It is the most damaging disease on<br />

oats and has probably been a pathogen since oats<br />

were first cultivated (Simons, 1985). In addition <strong>to</strong><br />

oats, numerous other grasses are infected, including<br />

important forage grasses such as the perennial<br />

ryegrass, Lolium perenne, for which separate<br />

resistance breeding programmes have been<br />

initiated (Kimbeng, 1999). Heavily infected<br />

meadows can acquire a distinctly orange hue<br />

due <strong>to</strong> the urediniospores being produced in<br />

great abundance. There are no formae speciales<br />

other than f. sp. avenae on oats because the<br />

delimitation between strains on different grasses<br />

is not clear cut. Puccinia coronata is macrocyclic,<br />

alternating with Rhamnus (buckthorn) shrubs.<br />

As with P. graminis and P. triticina, a standardized<br />

nomenclature for identifying races of P. coronata<br />

f. sp. avenae has been proposed. This is based on<br />

a differential comprising 16 oat cultivars carrying<br />

single resistance (Pc) genes (Chong et al., 2000).<br />

In <strong>to</strong>tal, about 100 resistance genes are known,<br />

and resistance breeding is the main strategy <strong>to</strong><br />

control oat rust. Heavy infections of susceptible<br />

cultivars can result in <strong>to</strong>tal crop failure.<br />

Even well-known rust species can harbour<br />

surprises upon close inspection, and one of these<br />

was the description of a form of P. coronata<br />

infecting Bromus inermis as its principal host.<br />

Anikster et al. (2003) described that germinating<br />

teliospores of this form produced only two<br />

basidiospores, each of which had four instead<br />

of the usual two nuclei and was self-fertile, i.e.<br />

infection on Rhamnus cathartica gave rise <strong>to</strong> aecia<br />

without spermogonia. This fungus is therefore<br />

homothallic, like the correlated microcyclic<br />

P. mesnieriana (Anikster & Wahl, 1985) and probably<br />

the barley form of P. coronata. In contrast,<br />

P. coronata f. sp. avenae is heterothallic with a<br />

tetrapolar mating system, i.e. two mating type<br />

loci with two alleles each, instead of the usual<br />

bipolar system (Narisawa et al., 1994).<br />

22.4.5 The origin of cereal rusts<br />

It is generally assumed that the cereal rusts<br />

evolved at the centre of origin of the wild grasses<br />

which were the progeny of cultivated cereals.<br />

The argument is strengthened if the alternate<br />

host is native <strong>to</strong> the same area. This <strong>to</strong>pic has<br />

been discussed extensively by Anikster and Wahl<br />

(1979) and Wahl et al. (1984). Puccinia graminis is<br />

thought <strong>to</strong> have evolved with Berberis vulgaris,<br />

which is of Asian origin, moving westwards <strong>to</strong><br />

an area from Transcaucasia <strong>to</strong> the Western<br />

Mediterranean, which is the centre of origin of<br />

wheat and rye. Westward migration must have<br />

occurred early in the his<strong>to</strong>ry of agriculture<br />

because 3300-year-old P. graminis-infected cereal


PUCCINIA AND UROMYCES<br />

629<br />

Fig 22.13 Teliospores of rust fungi belonging <strong>to</strong> the Pucciniaceae sensu Dietel. (a) Puccinia coronata.Notethe‘crown’ofspine-like<br />

wall extensions at the teliospore apex. (b) Uromyces appendiculatus.The teliospore is one-celled. (c,d) Phragmidium mucronatum.<br />

(c) Teliospore freshly mounted in water. (d) Spore after about15 min in water.The base of the pedicel has broken, exuding a<br />

large amount of mucilage which has been made visible by replacing the water with dilute Indian ink. (e) Triphragmium ulmariae.<br />

The teliospore is three-celled. (f) Gymnosporangiumfuscum. Note the long teliospore stalk. (a,b,e) <strong>to</strong> same scale; (c,d) <strong>to</strong> same scale.<br />

remains have been found in Israel (Kislev, 1982).<br />

The centre of origin of P. triticina and its aecial<br />

host, Thalictrum, is probably also in the Western<br />

Mediterranean, as is that of Puccinia striiformis.<br />

Rhamnus and wild oats are common in Israel,<br />

as is P. mesnieriana, suggesting that this is where<br />

P. coronata may have its origin. Wild barley<br />

species, especially Hordeum spontaneum, are also<br />

very common in Israel and adjacent countries<br />

where barley has been cultivated since the<br />

dawn of agriculture. Since the alternate host of<br />

P. hordei, Ornithogalum, is also found there and is<br />

profusely infected, this region may be where<br />

P. hordei evolved.<br />

22.5 Puccinia and Uromyces<br />

The genus Puccinia is by far the largest among<br />

the Uredinales, comprising about 4000 species<br />

(Kirk et al., 2001). The most readily recognized<br />

characteristic is the teliospore which usually<br />

contains two darkly pigmented cells borne on<br />

a thin hyaline stalk. Puccinia spp. are pathogenic<br />

mainly on Angiosperms, and especially on<br />

grasses as principal hosts. The genus Uromyces<br />

(about 800 spp.) is similar in appearance <strong>to</strong><br />

Puccinia, differing mainly in its teliospore, which<br />

is commonly one-celled (Fig. 22.13b). Several<br />

Puccinia spp. also produce a certain proportion<br />

of one-celled teliospores (called mesospores) in<br />

their telia, and it is therefore easy <strong>to</strong> imagine how<br />

the one-celled teliospore of Uromyces could have<br />

evolved. DNA sequencing data support a close<br />

relationship between the genera Puccinia and<br />

Uromyces (Maier et al., 2003).<br />

22.5.1 Other common Puccinia spp.<br />

Puccinia sorghi was so named because Schweinitz,<br />

when he first described it, erroneously believed


630 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

that he was dealing with infected Sorghum<br />

material. It is, however, a pathogen of maize<br />

and can be encountered wherever that crop is<br />

grown. Since maize is native <strong>to</strong> Central America,<br />

the pathogen probably also originated there.<br />

In Europe, crop losses are not generally sufficiently<br />

severe <strong>to</strong> warrant control measures<br />

(Smith et al., 1988), but elsewhere breeding programmes<br />

are in operation and fungicides are<br />

also occasionally used (Hooker, 1985). The fungus<br />

produces numerous chocolate-brown uredinia<br />

on the leaves of grain and fodder maize. It is<br />

macrocyclic and heteroecious, alternating with<br />

Oxalis spp.<br />

Several macrocyclic Puccinia spp. not causing<br />

agriculturally significant disease are frequently<br />

encountered in the field. An example is P. caricina<br />

which has sedges (Carex spp.) as its principal host<br />

and produces spermogonia and aecia on<br />

Urtica (stinging nettle). The latter are associated<br />

with growth deformations (Plate 12b). Another<br />

ubiqui<strong>to</strong>us species is P. poarum, with grasses of<br />

the genus Poa as principal and Tussilago farfara<br />

(coltsfoot) as alternate hosts. This species is<br />

unusual in completing two life cycles per<br />

growing season, with aecia appearing in May<br />

and August (Wilson & Henderson, 1966). Mint<br />

rust (P. menthae) is au<strong>to</strong>ecious and macrocyclic.<br />

All the above species are abundant in Europe and<br />

on most other continents.<br />

Puccinia punctiformis is a systemic au<strong>to</strong>ecious<br />

rust attacking Cirsium arvense (creeping<br />

thistle). In spring, infected plants are clearly<br />

distinguishable by their yellowish appearance<br />

and appressed leaves, and by the strong sweet<br />

smell associated with numerous spermogonia<br />

which develop all over the infected shoots.<br />

These aromatic substances include common<br />

fragrance molecules such as benzaldehyde,<br />

2-phenylethanol, indole and phenylacetaldehyde<br />

(Connick & French, 1991), and it is possible <strong>to</strong><br />

smell infected thistles from a distance of several<br />

metres. All these substances are produced by<br />

a wide range of fungi and, with the exception<br />

of indole, also by the host plant during<br />

flowering. The infections develop from a dikaryotic<br />

mycelium overwintering systemically in<br />

the roots<strong>to</strong>ck. The haploid condition of the<br />

spermogonial tissue is probably established by<br />

de-dikaryotization as new shoots are infected in<br />

spring. Transfer of spermatia <strong>to</strong> compatible<br />

flexuous hyphae results in dikaryotization, and<br />

the next structures <strong>to</strong> form resemble uredinia<br />

with chocolate-brown spores. They are sometimes<br />

called uredinoid aecia. The urediniospores<br />

can infect healthy thistles and normal uredinia<br />

develop on these. Later in the season, teliospores<br />

develop (Buller, 1950).<br />

A yet more specialized way <strong>to</strong> attract insect<br />

pollina<strong>to</strong>rs has been developed by a complex of<br />

rusts attacking Brassicaceae (Arabis spp.). Infected<br />

plants produce the same fragrances as Cirsium<br />

infected by P. punctiformis, plus many more<br />

(Raguso & Roy, 1998). Further, infected plant<br />

organs are morphologically modified <strong>to</strong> form<br />

flower-like structures called pseudoflowers.<br />

Intriguingly, these do not resemble the flowers<br />

of the host, but those of taxonomically unrelated<br />

plant species such as buttercups (Ranunculus spp.),<br />

which often grow in the same habitats (Roy, 1994).<br />

The combination of scent, shape, colour and<br />

nectar in this floral mimicry is necessary <strong>to</strong><br />

attract a wide range of insects (Roy, 1994; Roy &<br />

Raguso, 1997). Several rust species can stimulate<br />

the production of such pseudoflowers on Arabis<br />

spp., including the macrocyclic heteroecious<br />

P. monoica which alternates with grasses, and<br />

correlated au<strong>to</strong>ecious and microcyclic species<br />

(Roy et al., 1998).<br />

22.5.2 Uromyces<br />

Common species of Uromyces are U. ficariae, a<br />

microcyclic species forming brown telia on<br />

Ranunculus ficaria, and U. dactylidis with urediniospores<br />

and teliospores on grasses (Dactylis,<br />

Festuca and Poa) and aecia on Ranunculus spp.<br />

Uromyces dianthi produces uredinia and telia on<br />

Dianthus (carnation) and other ornamental flowers,<br />

with Euphorbia spp. as the alternate host.<br />

However, the most important species of Uromyces<br />

are those infecting legumes (Fabaceae), and<br />

because of the ease with which legumes such as<br />

Pisum, Phaseolus and Medicago can be cultivated in<br />

the labora<strong>to</strong>ry, their rusts have been used for<br />

numerous studies of fundamental physiological<br />

aspects such as differentiation of infection<br />

structures and haus<strong>to</strong>rial functioning, which


OTHER MEMBERS OF THE PUCCINIACEAE<br />

631<br />

have been mentioned at the beginning of this<br />

chapter.<br />

Uromyces appendiculatus is a macrocyclic au<strong>to</strong>ecious<br />

rust on French bean (Phaseolus vulgaris) and<br />

many other leguminous plants. Several varieties<br />

have been distinguished. The fungus survives the<br />

winter as teliospores on plant debris or, in milder<br />

conditions, as urediniospores (McMillan et al.,<br />

2003). As one would expect of a situation in which<br />

a rust fungus undergoing sexual reproduction is<br />

being controlled by the breeding of resistant<br />

cultivars, numerous (more than 200) physiological<br />

races of U. appendiculatus are known. Control is<br />

by means of resistant cultivars or, if these fail,<br />

with protectant fungicides (maneb and chlorothalonil).<br />

Strobilurin-type compounds are also<br />

now used against U. appendiculatus. Since beans<br />

are a higher-value crop than cereals, fungicide<br />

applications are generally more profitable.<br />

Uromyces viciae-fabae has a similar life cycle <strong>to</strong><br />

that of U. appendiculatus, but it infects broad bean<br />

(Vicia faba) and numerous other plants. It is a<br />

cosmopolitan species but seems <strong>to</strong> cause lesser<br />

damage than U. appendiculatus.<br />

Uromyces pisi is heteroecious, producing spermogonia<br />

and aecia on Euphorbia cyparissias and<br />

uredinia and telia on various leguminous hosts,<br />

especially pea (Pisum sativum). Infections on the<br />

alternate host are of biological interest because<br />

the fungus infects systemically and causes its<br />

alternate host <strong>to</strong> produce pseudoflowers, at the<br />

expense of its real flowers (Pfunder & Roy, 2000).<br />

The name U. pisi is now known <strong>to</strong> cover a<br />

complex of several biologically distinct species<br />

which are united by their alternate host but<br />

infect different members of the Fabaceae as<br />

principal hosts (Pfunder et al., 2001).<br />

22.6 Other members of the<br />

Pucciniaceae<br />

22.6.1 Phragmidium<br />

All Phragmidium spp. are au<strong>to</strong>ecious and confined<br />

<strong>to</strong> the Rosaceae. The two best-known species,<br />

both extremely common in Europe, are P. violaceum<br />

on leaves of the Rubus fruticosus species<br />

aggregate (bramble or blackberry; Plate 12c) and<br />

P. mucronatum on roses. Both are macrocyclic. The<br />

uredinial and telial stages are readily recognized<br />

from above as chlorotic or purple lesions, and<br />

the spores are formed on the underside of the<br />

host leaves. They are yellowish-orange (urediniospores)<br />

or violet <strong>to</strong> black (teliospores). The<br />

teliospores contain mostly 4 (P. violaceum) or<br />

6 8 (P. mucronatum) cells with very dark and<br />

rough walls, and are borne on a long stalk<br />

(pedicel) which becomes easily detached from the<br />

lesions. When teliospores are mounted in water,<br />

the base of the pedicel breaks and exudes a large<br />

amount of mucilage (Figs. 22.13c,d). In nature,<br />

this may facilitate the attachment of the teliospore<br />

<strong>to</strong> adjacent surfaces during overwintering<br />

(Ingold et al., 1981). The teliospores of P. mucronatum<br />

were among the first fungus spores illustrated<br />

by Robert Hooke in 1667 soon after the<br />

invention of the microscope (see Large, 1940).<br />

Phragmidium violaceum shows promise as a<br />

biocontrol agent against the spread of brambles<br />

which were introduced in<strong>to</strong> Australia and are<br />

spreading there in their typically uncontrollable<br />

manner (Mahr & Bruzzese, 1998). Phragmidium<br />

mucronatum is not <strong>to</strong>lerated by rose breeders or<br />

hobby gardeners; it is controlled by fungicide<br />

applications. Removal of fallen leaves in autumn<br />

can also help <strong>to</strong> control the disease.<br />

22.6.2 Gymnosporangium<br />

There are about 60 species of Gymnosporangium<br />

which produce their spermogonia and aecia<br />

on members of the Rosaceae, and telia on<br />

Cupressaceae. Uredinia are not normally<br />

formed. One of the most commonly encountered<br />

species is G. fuscum (¼ G. sabinae), the pear trellis<br />

rust, which alternates between Juniperus spp. of<br />

the sabina group (principal host) and pear trees.<br />

The dikaryotic mycelium can survive for many<br />

years in the principal host, and repeated production<br />

of telia is associated with a spindle-shaped<br />

swelling or canker of the infected trunk or twig.<br />

Spore-bearing telia are produced in the spring<br />

as horn-like outgrowths which greatly expand<br />

during rainfall due <strong>to</strong> swelling of the long teliospore<br />

stalks (Plate 12d; Fig. 22.13f). Teliospores at<br />

the surface of a swollen telial horn germinate<br />

<strong>to</strong> release basidiospores at the time of leaf


632 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

bud of the alternate host. Between late spring<br />

and autumn, infected pear leaves show bright<br />

orange-coloured lesions up <strong>to</strong> 1 cm in diameter<br />

(Plate 12e). These give rise <strong>to</strong> spermogonia at the<br />

upper leaf surface, and later <strong>to</strong> aecia at the lower<br />

surface. The aecia are greatly swollen and elongated.<br />

Eventually, they rupture at their sides,<br />

leaving a cap joined <strong>to</strong> the aecium by trellis-like<br />

threads (Plate 12f). The characteristic tube-like<br />

aecia of Gymnosporangium are referred <strong>to</strong> as<br />

roestelioid, after the generic name Roestelia<br />

previously given <strong>to</strong> such aecial stages. Aeciospores<br />

infect susceptible Juniperus spp. in<br />

the autumn (Borno & van der Kamp, 1975).<br />

Gymnosporangium fuscum can cause serious<br />

damage <strong>to</strong> pear trees because heavy infections<br />

reduce the pho<strong>to</strong>synthetic capacity of the trees,<br />

and yields and ultimately the trees themselves<br />

decline after severe successive infections. This<br />

species is very much on the increase in Europe<br />

and elsewhere, especially because of the planting<br />

of ornamental Juniperus spp. (e.g. Chinese juniper)<br />

and pear trees side-by-side in gardens. Whereas<br />

the pathogen needs <strong>to</strong> infect the alternate<br />

host afresh each year, it can survive almost<br />

indefinitely in its principal host once this has<br />

become infected. Several communities and<br />

countries have therefore launched eradication<br />

schemes <strong>to</strong> remove infected Juniperus trees.<br />

Since the basidiospores do not normally travel<br />

distances longer than about a mile, this approach<br />

can reduce the infection pressure <strong>to</strong> acceptable<br />

limits. Chemical control of G. fuscum on pear<br />

trees is possible but obviously not desirable<br />

(Ormrod et al., 1984).<br />

Several related species of Gymnosporangium are<br />

also commonly encountered, e.g. G. clavariiforme,<br />

which alternates between Juniperus communis<br />

(a plant not infected by G. fuscum) and Crataegus<br />

spp., or G. cornutum on J. communis and Sorbus<br />

aucuparia (mountain ash, rowan). Gymnosporangium<br />

juniperi-virginianae is the cause of North<br />

American cedar-apple rust, alternating between<br />

the ‘Eastern red cedar’ (Juniperus virginiana) and<br />

apple (Malus spp.). On the principal host, telia<br />

often develop from gall-like deformations called<br />

‘cedar-apples’. This species causes considerable<br />

economic damage in North America where it is<br />

native. It does not seem <strong>to</strong> be present in Europe as<br />

yet (Smith et al., 1988). It can be controlled by<br />

fungicide applications similar <strong>to</strong> those effective<br />

against apple scab caused by Venturia inaequalis<br />

(see p. 478).<br />

22.6.3 Hemileia vastatrix<br />

The phylogenetic position of H. vastatrix, the<br />

cause of coffee leaf rust, has not yet been<br />

resolved. It appears <strong>to</strong> be an ancient rust lineage<br />

possibly pre-dating the separation between<br />

Pucciniaceae and Melampsoraceae sensu Dietel<br />

(Wingfield et al., 2004). We assume that it belongs<br />

<strong>to</strong> the Pucciniaceae sensu Dietel because its<br />

teliospores are stalked. Coffee is one of the<br />

most valuable commodities on the world<br />

market, and H. vastatrix causes the most important<br />

disease of it. So far, only urediniospore,<br />

teliospore and basidiospore stages have been<br />

reported, and an alternate host has not been<br />

found. Therefore, it is likely that the disease is<br />

spread exclusively by urediniospores. Hemileia<br />

was so named because its urediniospores are<br />

unusual in being half smooth, with the other<br />

half of their surface spiny like the urediniospores<br />

of other rusts. Urediniospores of H. vastatrix<br />

infect coffee leaves in the typical rust fashion,<br />

i.e. they require free water for germination and<br />

form appressoria over s<strong>to</strong>mata. Resistance is<br />

expressed as a hypersensitive response after the<br />

formation of the first haus<strong>to</strong>rium (Silva et al.,<br />

2002). Unusual features include details of appressorium<br />

formation (Coutinho et al., 1993), the<br />

fact that infection occurs exclusively through<br />

s<strong>to</strong>mata located on the lower (abaxial) leaf<br />

surface, and the formation of urediniospores<br />

and teliospores through s<strong>to</strong>mata (Ward, 1882),<br />

i.e. typical rust pustules rupturing the epidermis<br />

are not formed. Another remarkable feature is<br />

that removal of the coffee berries, which are not<br />

themselves infected by H. vastatrix, drastically<br />

reduces the severity of the disease (Monaco,<br />

1977).<br />

The s<strong>to</strong>ry of coffee rust is fascinating and its<br />

first part has been <strong>to</strong>ld eloquently by Large<br />

(1940). The disease was discovered on cultivated<br />

coffee in Ceylon (now Sri Lanka) in about 1869.<br />

The origin of H. vastatrix seems obscure. Ceylon<br />

was then the major coffee-growing region for the


OTHER MEMBERS OF THE PUCCINIACEAE<br />

633<br />

British Empire in which coffee drinking was<br />

fashionable. Initial outbreaks of the rust were<br />

not taken seriously because severe infections,<br />

which led <strong>to</strong> the defoliation of trees, showed<br />

their effect only in subsequent seasons in the<br />

shape of a gradual debilitation of the coffee<br />

plants (see Brown et al., 1995). After the failure<br />

of several successive crops, H. Marshall Ward<br />

was sent <strong>to</strong> Ceylon in 1879 <strong>to</strong> investigate the<br />

problem, and he succeeded in elucidating the life<br />

cycle of H. vastatrix and the details of the<br />

infection process (Ward, 1882). Unfortunately,<br />

by that time it was <strong>to</strong>o late <strong>to</strong> save the coffee<br />

production of Ceylon and other South East Asian<br />

countries <strong>to</strong> which the rust had spread in the<br />

meantime. As a consequence of coffee rust, tea<br />

became the main crop of Ceylon, and tea<br />

drinking was promoted throughout the Empire<br />

(Schuman, 1991).<br />

The spread of coffee rust <strong>to</strong> all major coffeegrowing<br />

regions of the world (except Hawaii) is<br />

charted in Fig. 22.14. It is probably the combined<br />

result of human travel and natural long-distance<br />

transport. The much-dreaded jump from West<br />

Africa, where the disease arrived in the 1960s, <strong>to</strong><br />

Brazil probably occurred during a period between<br />

January and April in the late 1960s as a one-off<br />

transport of urediniospores by wind (Bowden<br />

et al., 1971). A similar chance event of wind-borne<br />

intercontinental rust spore traffic may have<br />

happened in June 1978 when the sugarcane rust<br />

(Puccinia melanocephala) arrived in the Dominican<br />

Republic, probably from Cameroon (Purdy et al.,<br />

1985; Brown & Hovmøller, 2002). Within coffee<br />

plantations, urediniospores of H. vastatrix may be<br />

spread both by wind and by rain splash.<br />

Coffee rust can be controlled by fungicide<br />

applications, with copper-containing compounds<br />

being the most useful even <strong>to</strong>day (Bock, 1962;<br />

Kushalappa & Eskes, 1989). Since the timing of<br />

fungicide application is crucial, disease forecast<br />

models <strong>to</strong> optimize fungicide applications are<br />

being developed. Resistance breeding is also<br />

promising. The main cultivated coffee plant,<br />

Coffea arabica, was clonally propagated by the<br />

Dutch in the late seventeenth century and is<br />

therefore genetically quite uniform across many<br />

coffee-growing areas, but the introduction of<br />

Fig 22.14 The journey of coffee rust (Hemileia vastatrix)<br />

aroundtheworld.Timepointsoffirstrecordsareindicated.<br />

Redrawn from Schuman (1991), with supplementary data from<br />

Monaco (1977) and Schieber and Zentmyer (1984).


634 UREDINIOMYCETES: UREDINALES (RUST FUNGI)<br />

resistance genes from wild coffee species can give<br />

rise <strong>to</strong> cultivars with stable partial resistance.<br />

Al<strong>to</strong>gether 9 major resistance genes and about 30<br />

relatively stable races of H. vastatrix are known<br />

(Monaco, 1977; Kushalappa & Eskes, 1989).<br />

22.7 Melampsoraceae<br />

22.7.1 Cronartium<br />

About 20 species of Cronartium are known. They<br />

are macrocyclic and heteroecious, alternating<br />

between various dicotyledonous plants as principal<br />

hosts and Pinus spp. as alternate hosts. The<br />

one-celled teliospores of Cronartium are produced<br />

in long columns projecting out of the telia. The<br />

teliospores are attached <strong>to</strong> each other but not <strong>to</strong><br />

their host tissue, and they are unstalked. In<br />

general, the mycelium on the alternate host is<br />

perennial whereas the principal host needs <strong>to</strong> be<br />

re-infected afresh each growing season. Reduced<br />

forms with aecia on Pinus spp. are named<br />

Peridermium (or Endocronartium). Their ability <strong>to</strong><br />

re-infect the Pinus host by means of aeciospores<br />

produced from aecia indicates that these are<br />

aecidioid uredinia (Ono, 2002). Cronartium and<br />

Peridermium spp. form a well-resolved monophyletic<br />

group of rust species also clustering<br />

<strong>to</strong>gether with their telial hosts (Vogler & Bruns,<br />

1998). Economic damage is generally caused<br />

mainly by the infection of Pinus spp. The<br />

symp<strong>to</strong>ms are similar between the various<br />

species considered below.<br />

Cronartium ribicola causes white pine blister<br />

rust, which alternates between Ribes spp.<br />

(currants, gooseberry etc.) and those Pinus spp.<br />

which produce their needles in groups of five<br />

(e.g. P. strobus). Infection of the alternate host<br />

proceeds by basidiospores germinating on the<br />

needles of Pinus spp. in autumn. During the<br />

following spring, infection spreads <strong>to</strong> the needle<br />

base and ultimately <strong>to</strong> the branches or main stem<br />

where lesions become visible as cankers. These<br />

produce spermogonia and aecia (Smith et al.,<br />

1988). Large cankers can girdle the entire trunk<br />

of trees, causing host death and widespread<br />

economic and ecological damage <strong>to</strong> Pinus forests<br />

especially in North America and Europe. In<br />

summer, aeciospores infect Ribes spp., where<br />

uredinial and telial lesions arise on the leaves.<br />

Severe infections can cause premature leaf<br />

abscission, but this stage of the disease is not<br />

economically as serious as that on Pinus. Itisat<br />

present impossible <strong>to</strong> control C. ribicola. Since<br />

its basidiospores cannot travel long distances, a<br />

massive attempt at eradicating the principal host<br />

was launched in North America. However, this<br />

campaign failed because Ribes spp. are <strong>to</strong>o<br />

abundant and re-grow easily from roots or<br />

seeds remaining in the soil (Maloy, 1997).<br />

<strong>Fungi</strong>cide treatment of infected trees was<br />

attempted by aerial sprays with cycloheximide<br />

(actidione), but this was ineffective (Dimond,<br />

1966; Maloy, 1997). More recently, attention<br />

has turned <strong>to</strong>wards the breeding of resistant<br />

cultivars. Although C. ribicola is probably of<br />

Asian origin and was introduced in<strong>to</strong> Europe<br />

in the eighteenth century and <strong>to</strong> North America<br />

around 1900, major gene resistance does exist<br />

among individual plants of Pinus spp. previously<br />

unexposed <strong>to</strong> the rust (Kinloch & Dupper, 2002).<br />

Resistance is often based on a gene-for-gene<br />

interaction whereby incompatibility commonly<br />

occurs as a hypersensitive response at the stage<br />

of needle infection (Jurgens et al., 2003). Resistance<br />

breeding holds promise because the populations<br />

of C. ribicola in Europe and North America<br />

are genetically quite uniform. The alternative <strong>to</strong><br />

resistance breeding is <strong>to</strong> give up P. strobus as a<br />

forestry tree, but the ecological damage <strong>to</strong> forests<br />

which regenerate by natural rejuvenation<br />

remains immense (Kinloch, 2003).<br />

Cronartium quercuum is the cause of fusiform<br />

rust especially on Pinus taeda and P. elliottii. These<br />

species were planted in the South-Eastern USA<br />

where C. quercuum originated on native, relatively<br />

resistant Pinus spp. The principal hosts are<br />

Quercus spp. The life cycle is similar <strong>to</strong> that of<br />

C. ribicola, as is the economic damage, with the<br />

exception that this species is still confined <strong>to</strong> the<br />

USA. The disease is now kept under control by<br />

the planting of rust-resistant cultivars (Powers &<br />

Kuhlman, 1997; Schmidt, 2003).<br />

Cronartium flaccidum causes resin <strong>to</strong>p disease<br />

on Scots pine (Pinus sylvestris), P. nigra, P. pinaster<br />

and other species in Europe (Smith et al., 1988).<br />

The principal hosts are various herbaceous


MELAMPSORACEAE<br />

635<br />

Fig 22.15 Section through a telial lesion of<br />

Melampsora euphorbiae on Euphorbia peplus.The<br />

teliospores are formed in a crust-like layer beneath<br />

the epidermis.<br />

plants. A similar disease is caused by the purely<br />

aecial (anamorphic) Peridermium pini. Both species<br />

are genetically identical, and Hantula et al. (2002)<br />

have recommended that they be considered as<br />

one species.<br />

22.7.2 Melampsora and Melampsoridium<br />

In Melampsora spp. the unicellular teliospores<br />

are sessile and often form a subepidermal crust<br />

(Fig. 22.15). Germination is by an external<br />

metabasidium of the usual type. The aecia lack<br />

peridia so that they are diffuse instead of cupshaped.<br />

Such diffuse aecia are called caeomata<br />

(singular caeoma). Melampsora lini var. lini is an<br />

au<strong>to</strong>ecious rust common on Linum catharticum,<br />

and M. lini var. liniperda infects cultivated flax.<br />

It was this fungus which Flor (1955, 1971) used<br />

for his pioneering work on the gene-for-gene<br />

hypothesis. Several Melampsora spp. (M. populnea,<br />

M. larici-populina, M. medusae, M. allii-populina)<br />

produce brightly coloured and extremely abundant<br />

uredinia and, later in the season, telia on<br />

Populus (poplar) trees (Plate 12g). Their alternate<br />

hosts are Larix, Abies, Picea, Pinus and Allium spp.<br />

(Smith et al., 1988). Some of these rusts seem <strong>to</strong><br />

differ very little other than in their aecial host,<br />

and much more work is required before an<br />

acceptable species concept is in place. Further,<br />

hybridization between different rusts can occur<br />

(Spiers & Hopcroft, 1994). There is an even more<br />

bewildering complex of Melampsora rusts on<br />

willows (Salix spp.), alternating with Allium,<br />

Larix, Ribes or orchid species (Smith et al., 1988;<br />

Pei et al., 1993). The most important species is<br />

M. epitea, which alternates with Larix and exists<br />

as numerous races. Since willow and poplar trees<br />

can be important crops in certain forest situations,<br />

efforts at resistance breeding are being<br />

made (Ramstedt, 1999).<br />

Melampsoridium betulinum causes a rust on<br />

birch as its uredinial and telial host, on which it<br />

is extremely common everywhere. Larix is the<br />

alternate host. In contrast <strong>to</strong> Melampsora spp.,<br />

aecia of Melampsoridium are surrounded by a<br />

peridium, and phylogenetic studies have also<br />

shown that the genera Melampsora and<br />

Melampsoridium are not particularly closely<br />

related (Maier et al., 2003).


23<br />

Ustilaginomycetes: smut fungi and their allies<br />

23.1 Ustilaginomycetes<br />

The Ustilaginomycetes are one of the four main<br />

classes of Basidiomycota and contain about 1500<br />

species (Kirk et al., 2001). In its present form as<br />

circumscribed by Begerow et al. (1997) and Bauer<br />

et al. (1997, 2001), this group is monophyletic.<br />

Hypha-producing Ustilaginomycetes are united<br />

by their lifestyle as ecologically obligate plant<br />

pathogens, often with an additional free-living<br />

(saprotrophic) yeast phase. They can be distinguished<br />

from the rust fungi in that haus<strong>to</strong>ria are<br />

either al<strong>to</strong>gether absent or, where present, take<br />

the shape of simple intracellular hyphae or<br />

hyphal extensions which invaginate the host<br />

plasmalemma but are not differentiated in<strong>to</strong> a<br />

narrow neck and a wider haus<strong>to</strong>rial body.<br />

Further, intracellular hyphae of Ustilaginomycetes<br />

usually secrete a thick sheath which is<br />

readily visible by transmission electron microscopy<br />

(see Figs. 23.6 and 23.17). The septa either<br />

lack perforations or contain simple pores or<br />

dolipores which are similar <strong>to</strong> those in the<br />

Urediniomycetes in lacking parenthesomes.<br />

True clamp connections are not usually found.<br />

The basidia of smut fungi produce numerous<br />

basidiospores whereas those of rust fungi usually<br />

produce only four.<br />

The class Ustilaginomycetes has been divided<br />

in<strong>to</strong> three subclasses by Begerow et al. (1997,<br />

2000). We shall consider representatives of two<br />

of these. The Ustilaginomycetidae (Section 23.2)<br />

are the most important plant-pathogenic<br />

Ustilaginomycetes, causing smut-like symp<strong>to</strong>ms.<br />

Typical members of the Exobasidiomycetidae<br />

(Section 23.4) cause other biotrophic diseases<br />

and are distinguished from the former by<br />

producing basidia directly from parasitic<br />

mycelium, not from teliospores. An exception<br />

is Tilletia, which is so similar <strong>to</strong> smuts in the<br />

Ustilaginomycetidae that we discuss it alongside<br />

them (p. 650). The third group, En<strong>to</strong>rrhizomycetidae,<br />

contains fungi which produce galls on the<br />

roots of Cyperaceae and Juncaceae (Vánky, 1994).<br />

Numerous Ustilaginomycetes live exclusively<br />

or predominantly as yeasts, and an example of<br />

this lifestyle is the order Malasseziales described<br />

on p. 670.<br />

<strong>Fungi</strong> causing smut-like disease symp<strong>to</strong>ms<br />

(Fig. 23.7, Plate 12h) have arisen at least<br />

twice within the Basidiomycota, namely in the<br />

Ustilaginomycetes and the Urediniomycetes.<br />

Urediniomyce<strong>to</strong>us smuts are found in the Microbotryales,<br />

and since these share many biological<br />

features with the ‘true’ smuts in the Ustilaginomycetes<br />

and have traditionally been studied by<br />

smut specialists, we shall consider the Microbotryales<br />

in this chapter (Section 23.3). Therefore,<br />

in our usage the term ‘smut fungus’ has a<br />

biological rather than a taxonomic meaning.<br />

23.2 The‘true’ smut fungi<br />

(Ustilaginomycetes)<br />

The word ‘smut’ describes the causal fungus or<br />

the symp<strong>to</strong>ms of a particular group of plant<br />

diseases in which loose masses (sori) of dark


THE ‘TRUE’ SMUT FUNGI (USTILAGINOMYCETES)<br />

637<br />

spores are produced in infected plant organs.<br />

Leaves, stems, flowers and seeds of grasses and<br />

other herbaceous plants are particularly frequently<br />

attacked. The term ‘bunt’ is sometimes<br />

used for smut fungi infecting the ovaries of their<br />

hosts, the seed becoming filled with teliospores<br />

in place of the embryo. Host tissues from which<br />

the spores are released often appear as if burnt<br />

or scorched, and this is why de Bary (1853)<br />

used the term ‘Brandpilze’ <strong>to</strong> describe the smut<br />

fungi. Various names have been given <strong>to</strong> these<br />

spores, e.g. teliospore, chlamydospore, brand<br />

spore, melanospore, us<strong>to</strong>spore or ustilospore.<br />

We prefer <strong>to</strong> call them teliospores because their<br />

function in the life cycle of smut fungi is<br />

equivalent <strong>to</strong> that of teliospores in the rust life<br />

cycle, i.e. they are the site of nuclear fusion and,<br />

on germination, give rise <strong>to</strong> the promycelium in<br />

which meiosis occurs.<br />

Several good monographs have been written<br />

about smut fungi, especially by Vánky (1987,<br />

1994). The surface ornamentation of teliospores<br />

(Figs. 23.2, 23.9) and their method of germination<br />

(Figs. 23.3, 23.4, 23.5, 23.10, 23.12) as well as the<br />

type of symp<strong>to</strong>ms and host range are important<br />

characters in identification.<br />

23.2.1 The life cycle of smut fungi<br />

The life cycle of smut fungi can be divided in<strong>to</strong><br />

two phases, a yeast-like monokaryotic (homokaryotic)<br />

form which grows saprotrophically but<br />

is unable <strong>to</strong> infect plants, and a predominantly<br />

dikaryotic (heterokaryotic) mycelial phase which<br />

is infectious on host plants but cannot grow<br />

saprotrophically (Fig. 23.1). The infectious dikaryotic<br />

mycelium is often systemic, showing interor<br />

intra-cellular growth. Meristematic host<br />

tissues are particularly densely colonized.<br />

Usually there are no specialized haus<strong>to</strong>ria, but<br />

hyphae may enter or even grow through host<br />

cells, invaginating the host plasmalemma.<br />

Intracellular hyphae are surrounded by a partial<br />

or complete sheath (Fig. 23.6) which is formed<br />

partly by the secre<strong>to</strong>ry activity of the pathogen<br />

and partly by the contributions from the infected<br />

host cell (Bauer et al., 1997). During sorus<br />

development, the dikaryotic hyphae proliferate<br />

and mass <strong>to</strong>gether in intercellular spaces, often<br />

destroying the softer internal host tissues but<br />

remaining enclosed by the host epidermis. There<br />

are no phialides or other specialized sporeproducing<br />

structures, but most hyphal segments<br />

can become converted <strong>to</strong> spores. The sporogenous<br />

hyphae are composed of binucleate cells.<br />

After nuclear fusion, the walls thicken and<br />

gelatinize, and the cells fragment. They then<br />

enlarge and become globose, and finally a thick<br />

wall is laid down (Snetselaar & Mims, 1994;<br />

Banuett & Herskowitz, 1996). The gelatinous<br />

matrix disappears at maturity. The ripe, uninucleate,<br />

diploid teliospores have thick, usually<br />

dark walls which may be smooth or ornamented<br />

by spines or reticulations (Vánky, 1987, 1994;<br />

Piepenbring et al., 1998a,b). In some genera, e.g.<br />

Urocystis, a central fertile cell is surrounded by a<br />

group of sterile cells.<br />

The teliospores are commonly dispersed by<br />

wind or become attached <strong>to</strong> the surface of seeds.<br />

Teliospores of many smut fungi germinate in a<br />

similar way <strong>to</strong> those of rust fungi by forming a<br />

septate promycelium. The teliospore is therefore<br />

the probasidium, with the promycelium being<br />

the metabasidium. If this is septate, it is a<br />

phragmobasidium. Teliospore germination can<br />

proceed in several different ways which are<br />

shown in subsequent sections. Typically, numerous<br />

haploid basidiospores (often called sporidia)<br />

are produced by direct budding from segments<br />

of the promycelium, and these are unable <strong>to</strong><br />

re-infect the host. Instead, they germinate by<br />

further budding <strong>to</strong> form elongated yeast cells<br />

which are capable of prolonged saprotrophic<br />

growth. Compatible yeast cells meeting on the<br />

surface of the host plant conjugate, and this<br />

initiates the dikaryotic hyphal stage which is<br />

able <strong>to</strong> infect the host plant.<br />

23.2.2 Mating systems<br />

Many smut fungi are heterothallic with a unifac<strong>to</strong>rial<br />

(bipolar) mating system (i.e. one mating<br />

type locus with two alleles), but Ustilago maydis<br />

and a few other species are tetrapolar, with two<br />

mating type loci located on different chromosomes.<br />

The a locus has two idiomorphs and<br />

controls fusion of sporidia which is driven by<br />

the exchange of mating pheromones between


638 USTILAGINOMYCETES: SMUT FUNGI AND THEIR ALLIES<br />

Fig 23.1 The life cycle of Ustilago maydis, diagrammatic and not <strong>to</strong> scale.Teliospores are formed in galls developing from dikaryotic<br />

hyphae. Karyogamy (K) occurs in the teliospore, followed by meiosis (M) and germination <strong>to</strong> form a three-septate promycelium.<br />

Each segment buds off numerous haploid sporidia. Sporidia bud in a yeast-like manner and are capable of saprotrophic growth but<br />

unable <strong>to</strong> infect maize plants.When yeast cells or sporidia of compatible mating type meet, conjugation tubes are formed, and their<br />

fusion (plasmogamy, P) gives rise <strong>to</strong> a dikaryotic hypha which is infectious on maize.Open and closed circles represent haploid nuclei<br />

of opposite mating type; the diploid nucleus is drawn larger and half-filled.<br />

compatible cells. The multiallelic b locus controls<br />

growth of the dikaryon, as well as its ability <strong>to</strong><br />

infect the host and <strong>to</strong> complete sexual development<br />

(see p. 643). In the bipolar species U. hordei,<br />

the genetic control of fusion and dikaryon growth<br />

is very similar <strong>to</strong> that in U. maydis, with the<br />

important differences that there are only<br />

two instead of multiple b alleles and that the<br />

a and b loci are tightly linked on the same<br />

chromosome (Lee et al., 1999). Genetic recombination<br />

during meiosis is suppressed in the region<br />

containing the a and b loci. This region is<br />

substantially larger than the mating type genes<br />

themselves, making up about one-sixth of the<br />

chromosome on which it resides. Hence, this<br />

chromosome is regarded as a primitive form of<br />

the sex chromosome as found, for example,<br />

in mammalian organisms (Fraser & Heitman,<br />

2004). Mammalian sex chromosomes are recognizably<br />

different when they are condensed at<br />

mi<strong>to</strong>sis, the X chromosome being very much<br />

larger than the Y chromosome. Intriguingly,<br />

measurable size differences have also been<br />

found between the two sex chromosomes of the<br />

urediniomyce<strong>to</strong>us smut, Microbotryum violaceum<br />

(Hood, 2002).<br />

The smut fungi are dimorphic, producing<br />

both yeast cells and true hyphae. This feature has<br />

aroused the interest of cell biologists, and<br />

especially Ustilago maydis is being used extensively<br />

as a <strong>to</strong>ol <strong>to</strong> examine fundamental aspects<br />

of eukaryotic biology (see Fig. 23.8).


THE ‘TRUE’ SMUT FUNGI (USTILAGINOMYCETES)<br />

639<br />

entire infected host organs or galls may be<br />

dispersed, and teliospores or mycelium of smut<br />

fungi may be borne on or in seeds.<br />

The most common dispersal route in the<br />

smut fungi is by wind-blown teliospores, and<br />

the transatlantic wind-borne passage of African<br />

sugarcane smut, Sporisorium scitamineum, has<br />

been proposed as the most likely route of entry<br />

in<strong>to</strong> the Caribbean. Teliospores of many smuts<br />

may adhere <strong>to</strong> the seeds of their host plants<br />

and germinate with them. Certain species causing<br />

loose smuts of cereals (see p. 639) systemically<br />

infect the living embryo of the cereal grain.<br />

Grains contaminated by external spore dusts or<br />

systemic infections can be spread by human<br />

transport. Insect dispersal is important in the<br />

anther smut, Microbotryum violaceum, which is<br />

related <strong>to</strong> the rust fungi (see p. 652).<br />

Fig 23.2 Teliospore surfaces of Ustilago spp. (a) Smooth<br />

surface in U. hordei. (b) Spiny surface in U. nuda.From<br />

Va¤nky (1994); original prints kindly provided by K.Va¤nky.<br />

23.2.3 Teliospore release and dispersal<br />

Piepenbring et al. (1998c) have written an<br />

excellent review of the teliospore as a dispersal<br />

unit in a wide range of smut fungi, and we can<br />

only summarize a few salient features here.<br />

Teliospores of most species are produced in a<br />

sorus enclosed by a thin layer of host tissue. They<br />

are generally released dry. Spore release is often<br />

by the simple rupture of the host epidermis<br />

surrounding the sorus, possibly as a result of<br />

pressure exerted by the expanding teliospores.<br />

The host epidermis may aid in the release of<br />

teliospores, e.g. if it is hit by water drops. The<br />

mechanism in this case is similar <strong>to</strong> that found<br />

in puffballs (see p. 578). Gusts of wind which<br />

shake infected host organs may also be effective<br />

in releasing smut spores over time. Alternatively,<br />

23.2.4 Ustilago species on<br />

monocotyledonous hosts<br />

Several Ustilago spp. cause diseases on grasses<br />

and cereals. Sori of the agriculturally most<br />

important species, U. hordei, U. nuda, U. tritici<br />

and U. avenae, are produced in place of the<br />

developing seeds. Other Ustilago spp. affect the<br />

leaves of their hosts, e.g. U. filiformis (formerly<br />

U. longissima), which causes leaf stripe smut<br />

on Glyceria spp. Ustilago maydis infects both<br />

vegetative and reproductive organs of its host<br />

(Plate 12h), causing gall-like deformations. This<br />

species is considered in detail on pp. 643 647.<br />

Many Ustilago spp. have also been described from<br />

dicotyledonous hosts, but it now seems that all<br />

of them belong <strong>to</strong> Microbotryum (Vánky, 1998,<br />

1999; Almaraz et al., 2002; see p. 652).<br />

Phylogenetic studies have shown U. hordei,<br />

U. nuda, U. tritici and U. avenae <strong>to</strong> be closely<br />

related <strong>to</strong> each other, <strong>to</strong> the point where it<br />

becomes difficult <strong>to</strong> distinguish individual<br />

species by DNA sequences and microscopy. They<br />

are regarded as the core species of Ustilago (S<strong>to</strong>ll<br />

et al., 2003). Since these species can hybridize<br />

with each other, it has been proposed that they<br />

should be merged in<strong>to</strong> one taxon, U. segetum<br />

(see Bakkeren et al., 2000). However, because<br />

the existing names are so well-established and<br />

because well-known plant diseases are caused by


640 USTILAGINOMYCETES: SMUT FUNGI AND THEIR ALLIES<br />

Fig 23.3 Teliospore germination<br />

in Ustilago avenae. (a) Germinating<br />

teliospore showing the four-celled<br />

promycelium, each cell of which<br />

is producing sporidia. Budding of<br />

detached sporidia is also shown.<br />

(b) Germinating teliospore showing<br />

fusion of the two terminal cells<br />

<strong>to</strong> initiate a dikaryon. (c) Fusion<br />

of conjugation tubes from two<br />

basidiospores <strong>to</strong> initiate a dikaryon.<br />

these cereal smuts, we prefer <strong>to</strong> keep their<br />

original names for the time being. The corn<br />

smut pathogen, U. maydis, is not closely related<br />

<strong>to</strong> the core Ustilago spp. and instead occupies an<br />

intermediate position between Ustilago and<br />

Sporisorium (S<strong>to</strong>ll et al., 2003). In <strong>to</strong>tal, there are<br />

about 230 Ustilago spp. and 190 Sporisorium spp.<br />

(Kirk et al., 2001), although such numbers<br />

obviously vary with the species concept adopted.<br />

The surface of teliospores is an important aid<br />

in identification. Ustilago hordei has a smooth<br />

teliospore surface whereas most other cereal<br />

smuts have spiny surfaces (Figs. 23.2a,b). Huang<br />

and Nielsen (1984) have shown that this difference<br />

between smooth and spiny surfaces is due<br />

<strong>to</strong> only two genes. Many of those Ustilago spp.<br />

now grouped with Microbotryum have teliospore<br />

surfaces with conspicuous reticulations or<br />

striations. There are also variations in the<br />

processes of teliospore germination in Ustilago<br />

(Ingold, 1983c; Vánky, 1994). The classical pattern<br />

is shown by U. avenae (Fig. 23.3), U. hordei and<br />

U. maydis in which the promycelium is threeseptate.<br />

All four cells give rise <strong>to</strong> sporidia, and<br />

compatible sporidia fuse following pheromone<br />

stimulation (see p. 643). When attached <strong>to</strong> a host<br />

plant surface, teliospores of U. avenae germinate<br />

in a different way; instead of producing sporidia,<br />

adjacent compatible cells of the promycelium<br />

fuse directly (Fig. 23.3b; Vánky, 1994). In U. nuda<br />

(Fig. 23.4), direct fusion of promycelium cells<br />

occurs both on agar and in nature. However,<br />

following synchronous division of the two nuclei<br />

in the fusion cell, septa are laid down such that<br />

a mosaic of both mono- and dikaryotic cells<br />

results. Compatible monokaryotic cells may fuse


THE ‘TRUE’ SMUT FUNGI (USTILAGINOMYCETES)<br />

641<br />

Fig 23.4 Teliospore germination in<br />

Ustilago nuda. (a) Three teliospores<br />

20 h after germination, showing<br />

various stages of development of<br />

the promycelium.The arrows<br />

indicate points where cell fusion<br />

has occurred. (b) A later stage<br />

showing the extension of<br />

mycelium from the fusion cells.<br />

(c) Two-day-old germinating<br />

teliospore showing repeated cell<br />

fusions (arrows).<br />

Fig 23.5 Teliospore germination in Ustilago<br />

filiformis. (a) Teliospore germination by successive<br />

development of sporidia. Note the absence of an<br />

extended promycelium.The first-formed sporidia<br />

are short. (b) Sporidium showing budding.<br />

(c) Sporidia conjugating. A dikaryotic mycelium<br />

arises following conjugation.


642 USTILAGINOMYCETES: SMUT FUNGI AND THEIR ALLIES<br />

Fig 23.6 TEM of Ustilago hordei infecting<br />

a compatible barley cultivar.The host cell<br />

consists of a large vacuole (V) surrounded<br />

by a thin layer of cortical cy<strong>to</strong>plasm.<br />

The intracellular fungal hypha (F) has<br />

penetrated the plant cell wall (PW) and<br />

invaginated the host plasmalemma (single<br />

arrow); note the thick electron-dense<br />

sheath (Sh) between the hyphal wall and<br />

the host plasmalemma.The double arrow<br />

points <strong>to</strong> vesicles in the intercellular space<br />

(IS). Pho<strong>to</strong>micrograph taken by G. Hu;<br />

reprinted from Hu et al.(2003),with<br />

permission from Elsevier.<br />

Fig 23.7 Loose smuts. (a) Infection of Ustilago avenae on oats.<br />

All seeds on the affected inflorescence have been replaced by<br />

teliospore sori. (b) Ustilago tritici causing loose smut on wheat.<br />

A healthy wheat ear is shown on the left.<br />

with each other again at a later stage (Nielsen,<br />

1988). In U. tritici there are also no sporidia,<br />

but here each uninucleate cell of the septate<br />

promycelium gives rise <strong>to</strong> a germ tube, and the<br />

dikaryophase is established by fusion of these<br />

germ tubes (Malik, 1974). Ustilago filiformis has<br />

teliospores which, on germination, do not<br />

produce an obvious promycelium, but merely a<br />

short tube from which sporidia are budded off<br />

successively (Fig. 23.5). All of the above fusion<br />

events ultimately result in a dikaryotic hypha<br />

which infects the host plant.<br />

In contrast <strong>to</strong> U. maydis, many small-grain<br />

cereal smuts entertain a gene-for-gene relationship<br />

with their hosts. In incompatible interactions,<br />

the infection hypha is encased in an<br />

unusually thick sheath as soon as it penetrates<br />

the first epidermal cell. This is closely followed<br />

by the death of the infected cell through a<br />

hypersensitive response (Hu et al., 2003). The<br />

breeding of resistant cultivars can be an<br />

effective means <strong>to</strong> control these pathogens<br />

(Smith et al., 1988). As we have already seen in<br />

the rust fungi (p. 625), the evolution of numerous<br />

races of smut fungi causes difficulties in<br />

cultivating cereals carrying major resistance<br />

genes. The spread of these races can be moni<strong>to</strong>red<br />

in time and space by screening field<br />

isolates against a differential of cereal cultivars<br />

(Menzies et al., 2003). Distinct smut races show a<br />

high tendency <strong>to</strong> hybridize under natural conditions,<br />

so that new races can arise rapidly<br />

(Thomas, 1984).


THE ‘TRUE’ SMUT FUNGI (USTILAGINOMYCETES)<br />

643<br />

Ustilago hordei<br />

This is the type species of Ustilago, causing<br />

covered smut of barley and oats. Crop losses<br />

are usually less than 1% in well-managed<br />

agrosystems (Thomas & Menzies, 1997). The<br />

infection cycle has been described in detail by<br />

Hu et al. (2002). The term ‘covered smut’ implies<br />

that the teliospores produced in the cereal grains<br />

replace the internal tissues but remain covered<br />

by the outer layer (pericarp) of the grains and<br />

are released only during threshing. Teliospores<br />

attached <strong>to</strong> the outer surface of seeds will<br />

germinate at the time of host germination.<br />

The fusion of compatible monokaryotic yeast<br />

cells results in a dikaryotic hypha which infects<br />

the young seedling. The epidermal cells are<br />

penetrated directly from slightly swollen<br />

hyphal tips, which may be regarded as rudimentary<br />

appressoria. There then follows an extended<br />

phase of systemic growth without outward<br />

symp<strong>to</strong>ms of infection. Colonization of the host<br />

is chiefly intracellular but also intercellular.<br />

Differentiated haus<strong>to</strong>ria are not produced.<br />

Instead, hyphae grow through host cells, invaginating<br />

the plasmalemma in the process. A thick<br />

sheath is deposited between the plasma membranes<br />

of the host and the pathogen (Fig. 23.6).<br />

By the time the mycelium reaches the apical<br />

meristem (about 40 55 days post infection),<br />

the host has usually formed the inflorescence<br />

initials, and massive proliferation of the mycelium<br />

occurs from that stage onwards. The<br />

developing spike tissue becomes filled with<br />

hyphae which then branch profusely and form<br />

teliospores. This infection sequence of a prolonged<br />

symp<strong>to</strong>mless colonization followed by<br />

symp<strong>to</strong>m development in the reproductive structures<br />

of the host is typical of the small-grain<br />

cereal smuts but differs from the infection of<br />

maize by U. maydis (see below).<br />

Ustilago avenae, U. nuda and U. tritici<br />

These species cause loose smut of their cereal<br />

hosts, i.e. the kernels are replaced by sori<br />

producing teliospores, with entire glumes<br />

usually destroyed. Ustilago avenae infects oats<br />

(Fig. 23.7a) and false oat-grass, Arrhenatherum<br />

elatius. There are numerous races which are<br />

specific <strong>to</strong> different oat cultivars. The disease<br />

cycle is complex. Teliospores have been shown <strong>to</strong><br />

survive for 13 years and thus the fungus can<br />

infect seedlings from teliospores dusted on<strong>to</strong> the<br />

outer surface of seeds. Additionally, it is capable<br />

of limited systemic infection of the seed pericarp<br />

which is initiated during flowering (Neergaard,<br />

1977), and such infections are genuinely seedborne.<br />

Either way, the fungus systemically<br />

colonizes the growing host plant and proliferates<br />

during flowering <strong>to</strong> produce a crop of teliospores.<br />

Systemically infected hosts flower slightly<br />

earlier than healthy plants, and they release<br />

their teliospores at the time when healthy plants<br />

flower. Teliospores released at that point can<br />

germinate on healthy flowers (Mills, 1967),<br />

leading <strong>to</strong> systemic infections which may be<br />

carried over <strong>to</strong> the next growing season within<br />

viable seeds.<br />

Ustilago nuda and U. tritici cause loose smut on<br />

barley and wheat, respectively (Fig. 23.7b). Their<br />

disease cycles are very similar <strong>to</strong> each other.<br />

Systemic infection of the host plant gives rise<br />

<strong>to</strong> smutted heads and, as in the case of U. avenae,<br />

flowering of smutted plants occurs slightly<br />

earlier than that of healthy plants, so that the<br />

disease can spread <strong>to</strong> uninfected flowers by<br />

means of teliospores. The normal entry point of<br />

dikaryotic mycelium was formerly thought <strong>to</strong> be<br />

the stigma of healthy flowers, but Batts (1955)<br />

showed that it is, in fact, the young tissue at<br />

the base of the ovary. The mycelium survives<br />

systemically in infected embryos. In spring,<br />

systemic infections spread when these embryos<br />

germinate. In contrast <strong>to</strong> U. avenae, teliospores of<br />

U. nuda and U. tritici are short-lived, rarely surviving<br />

for more than a few days under normal<br />

conditions. Hence, infection from teliospore dust<br />

on the surface of seeds does not seem <strong>to</strong> be an<br />

important infection route.<br />

23.2.5 Ustilago maydis<br />

This species is the cause of corn smut and is by<br />

far the most thoroughly researched member of<br />

the Ustilaginomycetes (reviewed by Kahmann<br />

et al., 2000). One of the early research highlights<br />

using U. maydis was on mi<strong>to</strong>tic recombination,<br />

leading <strong>to</strong> the development of the ‘Holliday<br />

model’ <strong>to</strong> explain the exchange of DNA strands


644 USTILAGINOMYCETES: SMUT FUNGI AND THEIR ALLIES<br />

at ‘Holliday junctions’ between paired DNA<br />

double helices (see p. 317; Holliday, 1962, 1964).<br />

Areas of ongoing research on U. maydis are<br />

described in subsequent sections. Work on<br />

U. maydis has greatly benefited from the ease<br />

with which the haploid phase of this fungus can<br />

be grown in culture and transformed by molecular<br />

biology <strong>to</strong>ols. The complete genome of this<br />

fungus has now been sequenced, and this should<br />

stimulate further research.<br />

The teliospores of U. maydis are long-lived and<br />

can survive in the environment (e.g. the soil) for<br />

several years. Dikaryotic hyphae resulting from<br />

the fusion of haploid sporidia can infect any<br />

above-ground tissue of its host (wild and cultivated<br />

Zea mays) by entry through s<strong>to</strong>mata or<br />

direct, appressorium-mediated penetration of the<br />

epidermis (Snetselaar & Mims, 1992; Banuett &<br />

Herskowitz, 1996). Infected host organs become<br />

strongly hypertrophied <strong>to</strong> form galls. This is<br />

in contrast <strong>to</strong> the small-grain cereal smuts<br />

described above, in which symp<strong>to</strong>m development<br />

is confined <strong>to</strong> the developing seeds of the host,<br />

and infection usually occurs at the seedling stage<br />

followed by a prolonged symp<strong>to</strong>mless phase. In<br />

U. maydis, the entire process from dikaryon<br />

formation on the plant surface <strong>to</strong> the release of<br />

mature teliospores may take as little as 2 weeks<br />

(Banuett & Herskowitz, 1996). In the field, infections<br />

by U. maydis are most commonly observed<br />

on the cobs presumably because the stigmata<br />

with their thin epidermal layer are most readily<br />

penetrated by the fungus (Snetselaar & Mims,<br />

1993). As a result of infection, the developing<br />

seeds on the corn cob become replaced by<br />

gall-like outgrowths (‘tumours’) which measure<br />

about 1 5 cm in diameter (Plate 12h). Although<br />

not generally welcomed by farmers, such infected<br />

cobs are prized as a delicacy in the Mexican<br />

cuisine (Pataky & Chandler, 2003).<br />

Monokaryotic yeast cells of U. maydis synthesize<br />

auxins, especially indole-3-acetic acid, in<br />

pure culture (see Basse et al., 1996), and infected<br />

hypertrophied host tissue also shows elevated<br />

concentrations of this plant growth hormone. It<br />

therefore seems likely that the production of<br />

growth hormones by U. maydis in planta contributes<br />

<strong>to</strong> the development of the striking disease<br />

symp<strong>to</strong>ms, although this has not yet been<br />

formally proven. The mycelium of U. maydis<br />

ramifies in these hypertrophied tissues, followed<br />

by hyphal fragmentation and production of teliospores.<br />

Although the dikaryotic phase is obligately<br />

biotrophic in nature, U. maydis can be<br />

stimulated <strong>to</strong> complete its life cycle in artificial<br />

conditions if it is grown on living cell cultures of<br />

Zea mays separated from the dikaryotic mycelium<br />

by a membrane permitting the diffusion of<br />

metabolites (Ruiz-Herrera et al., 1999). Deviations<br />

from the life cycle as shown in Fig. 23.1 have<br />

been described by Kahmann et al. (2000).<br />

Mating and dikaryon establishment<br />

Ustilago maydis is heterothallic and has two<br />

mating type loci. A given dikaryon is fully<br />

pathogenic only if it contains different idiomorphs<br />

or alleles at both loci. Since it will<br />

produce haploid progeny of four genetically<br />

distinct types, the mating system is said <strong>to</strong><br />

be tetrapolar. Locus a has two idiomorphs,<br />

a1 and a2, each of which encodes a mating<br />

pheromone and the recep<strong>to</strong>r for the pheromone<br />

encoded by the opposite idiomorph. Two haploid<br />

cells will mate if they contain opposite idiomorphs<br />

at locus a, irrespective of their b alleles.<br />

Conjugation can be induced experimentally even<br />

between cells containing the same idiomorph if<br />

the matching pheromones are added. Mating is<br />

therefore purely driven by the pheromones<br />

which are peptides containing 13 (peptide a1)<br />

or 9 (peptide a2) amino acids derivatized with<br />

a lipid (farnesyl) side-chain (Spellig et al., 1994).<br />

A cell whose recep<strong>to</strong>r has bound the pheromone<br />

of the opposite mating type will arrest its<br />

cell cycle in the G2 position and undergo a<br />

morphogenetic change <strong>to</strong> produce a thin flexible<br />

conjugation tube which grows chemotropically<br />

<strong>to</strong>wards the pheromone source (Snetselaar et al.,<br />

1996).<br />

The ability <strong>to</strong> infect maize plants and <strong>to</strong><br />

complete the life cycle is tightly linked <strong>to</strong> the<br />

ability of the fusion cell <strong>to</strong> form dikaryotic<br />

hyphae. This is controlled by the b locus after<br />

conjugation of compatible cells. In contrast <strong>to</strong><br />

the a locus which has two idiomorphs with low<br />

sequence homology, the b locus has about 25<br />

alleles which are genetically relatively similar <strong>to</strong><br />

each other. Each allele encodes two proteins


THE ‘TRUE’ SMUT FUNGI (USTILAGINOMYCETES)<br />

645<br />

Fig 23.8 The roles of the a and b<br />

loci in the conjugation between<br />

monokaryotic yeast cells and<br />

the establishment of dikaryotic<br />

filamen<strong>to</strong>us growth in Ustilago<br />

maydis.Signallingeventsare<br />

represented by dotted lines;<br />

diffusion or transport of signalling<br />

molecules is indicated by dashed<br />

lines. Based on Bo«lker (2001).<br />

called bE and bW (Fig. 23.8), and the dikaryotic<br />

phenotype depends on the formation of a<br />

heterodimer of proteins transcribed from different<br />

alleles. For instance, bE1 and bW2 can<br />

dimerize but bE1 and bW1 cannot. The heterodimer<br />

is a transcription fac<strong>to</strong>r which triggers<br />

signalling cascades involved in maintaining<br />

dikaryon stability, hyphal growth and virulence<br />

on maize. At the same time, further mating<br />

reactions are suppressed.<br />

Signalling events involved in conjugation and<br />

the subsequent dimorphic switch from yeast-like<br />

monokaryotic <strong>to</strong> hyphal dikaryotic growth<br />

are being extensively examined and have been<br />

reviewed, for example, by Banuett (1995), Bölker<br />

(2001), Martínez-Espinoza et al. (2002) and<br />

Feldbrügge et al. (2004). Although it is impossible<br />

<strong>to</strong> discuss the details here, we can summarize<br />

that a cAMP and protein kinase A signalling<br />

chain is involved in maintaining growth by<br />

budding, whereas a MAP kinase cascade is crucial<br />

for the switch <strong>to</strong> hyphal growth and pathogenicity.<br />

These two branches seem <strong>to</strong> converge on<br />

a transcription fac<strong>to</strong>r called Prf1 which is subject<br />

<strong>to</strong> phosphorylation at different sites by protein<br />

kinase A and a MAP kinase (see Fig. 23.8).<br />

Of course, the cAMP/protein kinase A and MAP<br />

kinase signalling chains fulfil many other


646 USTILAGINOMYCETES: SMUT FUNGI AND THEIR ALLIES<br />

functions at subsequent developmental stages<br />

such as gall formation and teliospore production<br />

(Gold et al., 1997). For instance, the fuz7 gene<br />

product which encodes a MAP kinase kinase<br />

involved in the signalling chain shown in<br />

Fig. 23.8 is also involved in the process of<br />

hyphal fragmentation leading <strong>to</strong> teliospore production<br />

(Banuett & Herskowitz, 1996). Further,<br />

there is extensive cross-talk between the cAMP/<br />

protein kinase A and MAP kinase chains at<br />

different time-points in development.<br />

The cy<strong>to</strong>skele<strong>to</strong>n<br />

Much valuable work has been performed on<br />

U. maydis <strong>to</strong> assign various cellular transport<br />

phenomena <strong>to</strong> particular elements of the cy<strong>to</strong>skele<strong>to</strong>n<br />

and their associated mo<strong>to</strong>r proteins.<br />

Thus, microtubules have been implicated in<br />

the transport of nuclei, mi<strong>to</strong>chondria, vacuoles<br />

and the endoplasmic reticulum, as well as<br />

secre<strong>to</strong>ry vesicles in exocy<strong>to</strong>sis and endosomes<br />

in endocy<strong>to</strong>sis (Steinberg, 2000; Basse &<br />

Steinberg, 2004; Steinberg & Fuchs, 2004).<br />

Bidirectional movement can be achieved by<br />

kinesin mo<strong>to</strong>rs which move <strong>to</strong>wards the polymerizing<br />

end (plus end) of microtubules, and<br />

dynein which moves <strong>to</strong>wards the minus end.<br />

Actin cables and their myosin mo<strong>to</strong>rs are also<br />

involved in morphogenetic events and transport<br />

processes. Since the genome of U. maydis has been<br />

completely sequenced, the number of genes<br />

encoding myosin (3), dynein (1) and kinesin (10)<br />

is known (Basse & Steinberg, 2004) and further<br />

rapid progress on the role of the cy<strong>to</strong>skele<strong>to</strong>n<br />

in morphogenesis and cellular transport can be<br />

expected with U. maydis as an experimental<br />

organism.<br />

Mycoviruses and killer <strong>to</strong>xins in U. maydis<br />

Although mycoviruses are not uncommon in<br />

fungi, few of them have been investigated<br />

in detail. Two examples we have encountered in<br />

earlier chapters of this book are the virus-like<br />

particles in S. cerevisiae (p. 273) and related yeasts<br />

which contain double-stranded RNA (dsRNA)<br />

encoding killer <strong>to</strong>xins, and the hypovirulencecausing<br />

dsRNA viruses of Cryphonectria parasitica<br />

(p. 375) and Ophios<strong>to</strong>ma novo-ulmi (p. 366).<br />

Mycoviruses infecting Ustilago maydis are similar<br />

<strong>to</strong> those of S. cerevisiae in that they encode killer<br />

<strong>to</strong>xins. The first evidence of them was found<br />

when certain U. maydis strains killed sexually<br />

compatible strains upon anas<strong>to</strong>mosis in mating<br />

assays. The cy<strong>to</strong>plasmic inheritance of the killing<br />

trait, the proteinaceous nature of the <strong>to</strong>xin and<br />

the presence of virus particles in killer strains<br />

were quickly established (Hankin & Puhalla, 1971;<br />

Wood & Bozarth, 1973). There are three types<br />

of virus (P1, P4 and P6) each encoding its own<br />

killer protein (KP) <strong>to</strong>xin. Day (1981) showed that<br />

virus-infected U. maydis strains are common in<br />

field populations.<br />

A great deal is now known about viruses of<br />

U. maydis (see Magliani et al., 1997; Martínez-<br />

Espinoza et al., 2002). Their genomes are fragmented<br />

in<strong>to</strong> three size classes of dsRNA, whereby<br />

each size class can have several members.<br />

In <strong>to</strong>tal, there are six dsRNA fragments in P1,<br />

seven in P4 and five in P6, with one capsid able<br />

<strong>to</strong> accommodate either one H or one <strong>to</strong> several<br />

M chains (Bozarth et al., 1981). In all three<br />

viruses, a heavy (H) segment encodes the capsid<br />

protein and the replication machinery, and<br />

H segments are also essential for the maintenance<br />

of the medium-sized (M) and light (L)<br />

segments. The <strong>to</strong>xins are encoded by the<br />

M fragments, and their synthesis as prepropolypeptide<br />

chains followed by proteolytic<br />

cleavage and secretion of the mature <strong>to</strong>xins<br />

is similar <strong>to</strong> that of the S. cerevisiae killer<br />

<strong>to</strong>xins (p. 273). The function of the L segments<br />

is unknown at present.<br />

The modes of action of the three <strong>to</strong>xins seem<br />

<strong>to</strong> be diverse. The best-characterized is KP4 which<br />

is active as a monomer blocking certain types<br />

of Ca 2þ uptake channel. This activity can be<br />

observed in susceptible U. maydis strains as well<br />

as in mammalian cells, where it acts in a similar<br />

way <strong>to</strong> the black mamba snake venom, calciseptine<br />

(Gage et al., 2001, 2002). The KP1 and KP6<br />

<strong>to</strong>xins are released as two separate polypeptides<br />

after proteolytic cleavage, but in contrast <strong>to</strong> the<br />

yeast killer <strong>to</strong>xins these do not re-associate with<br />

each other by covalent (disulphide) bonds. Whilst<br />

the mode of action of KP1 is unknown, KP6 may<br />

form membrane pores which disrupt the ionic<br />

balance of the target cells (N. Li et al., 1999). The<br />

<strong>to</strong>xin-producing cell must obviously be resistant


THE ‘TRUE’ SMUT FUNGI (USTILAGINOMYCETES)<br />

647<br />

<strong>to</strong> its own <strong>to</strong>xin. In contrast <strong>to</strong> the S. cerevisiae<br />

system where the unprocessed <strong>to</strong>xin precursors<br />

bes<strong>to</strong>w resistance, in <strong>to</strong>xin-producing U. maydis<br />

strains resistance is encoded by nuclear genes.<br />

Resistance is specific, i.e. a KP1-producing strain<br />

is sensitive <strong>to</strong> KP4 and KP6. All three <strong>to</strong>xins are<br />

also effective against other members of the<br />

Ustilaginales (Koltin & Day, 1975).<br />

Attempts have been made <strong>to</strong> transform<br />

wheat plants with the KP4 <strong>to</strong>xin gene of the<br />

U. maydis virus P4 in order <strong>to</strong> engineer cultivars<br />

with resistance against Tilletia caries (Clausen<br />

et al., 2000), but no recent reports seem <strong>to</strong> have<br />

been published on this subject. Even if cultivars<br />

with good resistance in the field can be<br />

produced, it is unlikely that a transgenic crop<br />

plant expressing a calciseptine-like <strong>to</strong>xin will<br />

gain acceptance with regula<strong>to</strong>ry authorities or<br />

the general public.<br />

23.2.6 Tilletia<br />

There are about 125 species of Tilletia, all of which<br />

parasitize grasses (Poaceae). In economic terms,<br />

the most important pathogens are T. caries<br />

(¼ T. tritici) and T. indica on wheat, T. controversa<br />

on wheat and other cereals, and T. horrida on rice.<br />

Many other Tilletia spp. infect wild grasses (Vánky,<br />

1994). Teliospores of Tilletia spp. typically bear<br />

reticulate surface ornamentations (Fig. 23.9).<br />

They are long-lived in the soil and on the exterior<br />

of seeds. Infections are systemic and result in<br />

covered smut symp<strong>to</strong>ms in the seeds.<br />

past, but are much less frequent now because the<br />

incidence of common bunt has declined.<br />

Crushed sori have a fishy smell caused by the<br />

presence of trimethylamine. For this reason, the<br />

disease is also known as ‘stinking smut’, and<br />

flour made from contaminated grain is unfit for<br />

human consumption. Since teliospores are not<br />

readily released in the field, Piepenbring et al.<br />

(1998c) have suggested that T. caries may be an<br />

example of a fungus that has adapted <strong>to</strong> humans<br />

as a dispersal vec<strong>to</strong>r.<br />

This fungus was a major pathogen in the past<br />

and occupies a special place in the his<strong>to</strong>ry of<br />

plant pathology (Large, 1940; Ainsworth, 1981).<br />

In 1752, Mathieu Tillet, by careful experimentation,<br />

demonstrated that the common bunt<br />

disease was associated with dusting the seeds<br />

with bunt spores prior <strong>to</strong> sowing. He also<br />

reported that incidence of the disease was<br />

somewhat reduced by steeping the seeds in<br />

sea water and lime prior <strong>to</strong> sowing. In 1807,<br />

Bénédict Prévost observed the process of teliospore<br />

germination with his microscope, and he<br />

proposed that the bunt disease was caused by<br />

a living organism. Further, he discovered that<br />

teliospore germination was inhibited by copper<br />

salts, and that the treatment of wheat seeds with<br />

dilute copper sulphate solution was effective<br />

against infections of T. caries. The combination of<br />

copper and lime became known as the famous<br />

Bordeaux mixture only much later, in about<br />

Tilletia caries<br />

This species is cosmopolitan and causes ‘common<br />

bunt’, the best-known covered smut disease of<br />

wheat. The entire interior of the infected grain<br />

becomes converted <strong>to</strong> a greenish-brown teliospore<br />

sorus surrounded by the pericarp. Such sori<br />

are called ‘bunt balls’. The teliospores are not<br />

released until threshing, when they are dusted<br />

on<strong>to</strong> the surface of healthy grains. If heavily<br />

contaminated crops are processed, the teliospore<br />

concentrations in the air can be sufficiently high<br />

<strong>to</strong> cause dust explosions in mills or s<strong>to</strong>rage<br />

facilities. Respira<strong>to</strong>ry allergies among millers<br />

caused by T. caries were also common in the<br />

Fig 23.9 SEM of teliospores of Tilletia caries.FromVa¤nky<br />

(1994); original prints kindly provided by K.Va¤nky.


648 USTILAGINOMYCETES: SMUT FUNGI AND THEIR ALLIES<br />

Fig 23.10 Tilletia caries.<br />

(a) Germinating teliospore showing<br />

a non-septate promycelium and<br />

a crown of primary sporidia.<br />

(b) Two detached primary sporidia<br />

showing conjugation. A secondary<br />

sporidium has developed from<br />

one of the primary sporidia.<br />

(c) Primary sporidia attached<br />

<strong>to</strong> the promycelium showing<br />

conjugation.<br />

1882, when Alexis Millardet discovered its ability<br />

<strong>to</strong> control downy mildew of vines (see p. 119).<br />

The process of germination of T. caries teliospores<br />

is shown in Fig. 23.10. The diploid nucleus<br />

in the mature teliospore divides meiotically,<br />

and one or more mi<strong>to</strong>tic divisions follow so that<br />

8 or 16 nuclei are formed. The promycelium<br />

is often but not invariably aseptate, and from<br />

its tip narrow curved uninucleate primary<br />

sporidia arise, corresponding in number <strong>to</strong> the<br />

number of nuclei in the young promycelium<br />

(Fig. 23.10a). The primary sporidia mate in pairs<br />

by means of short conjugation tubes, often whilst<br />

still attached <strong>to</strong> the tip of the promycelium<br />

(Fig. 23.10c). Detached primary sporidia may also<br />

conjugate. During conjugation, a nucleus from<br />

one primary sporidium passes in<strong>to</strong> the other<br />

sporidium which therefore becomes binucleate.<br />

Each H-shaped pair of primary sporidia develops<br />

a single lateral sterigma on which a curved<br />

binucleate spore develops (Fig. 23.10b). This<br />

spore is projected actively from the sterigma<br />

using the surface-tension catapult mechanism,<br />

and it is sometimes referred <strong>to</strong> as the secondary<br />

sporidium. Because of its characteristic method<br />

of discharge, Buller and Vanterpool (1933) have<br />

interpreted this spore as a basidiospore. The<br />

secondary sporidium brings about infection of<br />

the host.<br />

Teliospores of T. caries (Fig. 23.9) are viable<br />

for up <strong>to</strong> 15 years and germinate along with the<br />

seeds if contaminated grain is sown. Secondary<br />

sporidia produce germ tubes which infect the<br />

coleoptiles of the seedlings. Infection is systemic,<br />

the mycelium growing through the tissues of the<br />

shoot, and by suitable techniques it is possible<br />

<strong>to</strong> isolate the dikaryotic mycelium from infected<br />

host tissues (Trione, 1972). Although infected<br />

plants may grow less vigorously than uninfected<br />

ones, they show no outward sign of<br />

infection until the ears are almost ripe.<br />

Tilletia controversa<br />

This species is closely related <strong>to</strong> T. caries and<br />

is most damaging on wheat, in addition <strong>to</strong><br />

infecting other cereals and grasses. The crucial<br />

difference is that teliospores of T. controversa<br />

germinate at much lower temperatures than<br />

those of T. caries. In consequence, T. controversa<br />

causes bunt mainly on winter wheat, infecting


THE ‘TRUE’ SMUT FUNGI (USTILAGINOMYCETES)<br />

649<br />

overwintering plants under snow cover from soilborne<br />

teliospores (Purdy et al., 1963). In contrast<br />

<strong>to</strong> other Tilletia spp., spores dusted on<strong>to</strong> the seed<br />

surface are unimportant as inoculum because<br />

germination of winter wheat seeds precedes<br />

teliospore germination by a few months. The<br />

disease caused by T. controversa is called ‘dwarf<br />

bunt’ because infected plants show stunted<br />

growth. Like T. caries, T. controversa has a genefor-gene<br />

relationship with its host, and breeding<br />

for resistance is an efficient control method<br />

(Fuentes-Dávila et al., 2002). The life cycle of<br />

T. controversa is similar <strong>to</strong> that of T. caries, except<br />

that T. controversa has a bipolar multiallelic<br />

mating system whereas that of T. caries is bipolar<br />

with only two mating type alleles.<br />

Tilletia controversa is cosmopolitan but is not in<br />

itself a serious pathogen. However, it has<br />

acquired no<strong>to</strong>riety by being used as the reason<br />

for establishing import restrictions on wheat<br />

imports from the United States by China during<br />

the 1970s <strong>to</strong> 1990s (Mathre, 1996).<br />

Tilletia indica<br />

The disease caused by T. indica is called Karnal<br />

bunt, named after the city in India where it was<br />

first described (Mitra, 1931). It is also called<br />

partial bunt because the teliospore sori often fill<br />

only part of the infected wheat grain, leaving the<br />

embryo unaffected. In contrast <strong>to</strong> T. caries and<br />

T. controversa, T. indica does not appear <strong>to</strong> grow<br />

systemically. Instead, teliospores germinate by<br />

producing numerous (up <strong>to</strong> 180) haploid primary<br />

needle-shaped sporidia from the tip of the<br />

aseptate promycelium. The primary sporidia<br />

germinate by monokaryotic haploid hyphae,<br />

and these in turn produce a further crop of<br />

haploid monokaryotic secondary sporidia. Two<br />

types may be produced, namely a repetition of<br />

the needle-shaped form or a sausage-shaped<br />

sporidium liberated actively by the surfacetension<br />

catapult mechanism. In contrast <strong>to</strong><br />

T. caries and T. controversa, H-shaped fusion cells<br />

are not formed and both types of secondary<br />

sporidium are monokaryotic and haploid.<br />

Secondary sporidia can germinate <strong>to</strong> produce a<br />

mycelium giving rise <strong>to</strong> further sporidia of either<br />

type (Dhaliwal, 1989). The actively liberated form<br />

enables the fungus <strong>to</strong> work its way up on the<br />

outside of the leaves of growing wheat plants<br />

until it reaches the flag leaf. There, fusion<br />

of compatible monokaryotic haploid hyphal<br />

segments may occur, with the resulting heterokaryotic<br />

mycelium causing infections of individual<br />

florets of the immature wheat ear (Goates,<br />

1988; Nagarajan et al., 1997).<br />

Occurrence and severity of Karnal bunt have<br />

increased in high-yielding crop systems, but the<br />

disease is not in itself serious, causing crop losses<br />

less than 1% per annum even in severe epidemics<br />

(Nagarajan et al., 1997). In addition, several<br />

resistant cultivars are available. Although bunt<br />

balls contain trimethylamine, flour made from<br />

crops with up <strong>to</strong> 4% infected grains is still fit for<br />

human consumption (Fuentes-Dávila et al, 2002).<br />

As in the case of T. controversa, the major threat<br />

posed by T. indica is a legal one. Countries as yet<br />

free from the disease may ban wheat imports<br />

from those affected by T. indica. For example, the<br />

United States imposed quarantine regulations<br />

<strong>to</strong> prevent the spread of Karnal bunt <strong>to</strong> North<br />

America, only <strong>to</strong> find bans imposed on US<br />

exports by other countries when T. indica was<br />

eventually discovered in Arizona in 1996 (Palm,<br />

1999). Rush et al. (2005) have given a fascinating<br />

account of the complex interactions between<br />

agriculture, politics, international trade and<br />

research in dealing with Karnal bunt in the<br />

United States, concluding that the threat posed<br />

by this disease was initially overstated.<br />

23.2.7 Urocystis<br />

The distinguishing feature of the genus Urocystis<br />

is that the teliospore consists of one or more<br />

melanized fertile cells surrounded by several<br />

sterile cells (Figs. 23.11, 23.12). This type of<br />

compound teliospore is called a spore ball.<br />

There are about 140 species (Vánky, 1994). An<br />

important species is U. tritici which causes leaf<br />

stripe-smut or flag smut on wheat in warm<br />

climates. This was formerly called U. agropyri, a<br />

name now applied in a more restricted sense <strong>to</strong><br />

forms on wild grasses such as Agropyron and<br />

Elymus spp. (Fig. 23.11). The fertile cell of a spore<br />

ball of Urocystis germinates by producing an<br />

aseptate promycelium which gives rise <strong>to</strong> about<br />

four primary sporidia (Fig. 23.12). These sporidia


650 USTILAGINOMYCETES: SMUT FUNGI AND THEIR ALLIES<br />

Fig 23.11 Teliospores of Urocystis agropyri.<br />

A dark-walled fertile cell is surrounded<br />

by several flattened hyaline sterile cells.<br />

Fig 23.12 Urocystis anemones. (a) Spore<br />

balls showing germination after 24 h<br />

incubation. A promycelium develops from<br />

a fertile cell, bearing a crown of three or<br />

four sporidia. (b) Development after 48 h.<br />

The sporidia have fused (arrow) and a<br />

septate infection hypha develops from the<br />

fusion cell.<br />

conjugate in pairs whilst still attached <strong>to</strong> the<br />

promycelium. The two haploid nuclei then<br />

migrate in<strong>to</strong> an infection hypha which develops<br />

from the conjugation tube (Thirumalachar &<br />

Dickson, 1949).<br />

Urocystis tritici infects its wheat host by<br />

appressorium-mediated penetration of the<br />

infection hypha through the host epidermis<br />

(Nelson & Durán, 1984). Infection occurs at<br />

the early seedling stage and is systemic. The outbreak<br />

of symp<strong>to</strong>ms is delayed, with the typical<br />

stripe-like sori breaking through the epidermis<br />

of several leaves after flowering of the host plant.<br />

Heavily infected plants may abort their tillers.


THE ‘TRUE’ SMUT FUNGI (USTILAGINOMYCETES)<br />

651<br />

Teliospores of U. tritici survive in the soil and<br />

attached <strong>to</strong> seeds (Fuentes-Dávila et al., 2002).<br />

Urocystis agropyri is common on wild grasses, and<br />

another common species is U. anemones which<br />

causes a leaf smut on Anemone and Ranunculus<br />

leaves.<br />

23.2.8 Control of smut diseases<br />

Control of loose and covered smuts presents very<br />

different problems. Whilst the surface of the<br />

grain is merely contaminated with the spores of<br />

covered smuts, in the case of loose smuts the ripe<br />

grain is already infected by a mycelium within the<br />

embryo. The control of covered smuts by means<br />

of fungicidal seed dressings is therefore simple,<br />

and it is standard practice for seed grains <strong>to</strong> be<br />

treated by seed merchants in this way. The first<br />

effective seed treatment was the steeping of seed<br />

grains in a dilute copper sulphate solution prior<br />

<strong>to</strong> sowing (Large, 1940). In the middle of the<br />

twentieth century, seed dressings containing<br />

organic mercury compounds were widely used,<br />

but these are now banned due <strong>to</strong> their high<br />

general <strong>to</strong>xicity. Instead, cocktails of fungicides<br />

are employed. In most countries with a welldeveloped<br />

agriculture, bunt of wheat is now a<br />

rare disease. For example, the incidence of bunt<br />

balls in seed samples sent <strong>to</strong> the Official Seed<br />

Testing Station at Cambridge fell from 12 33%<br />

in 1921 1925 <strong>to</strong> 0.2 0.3% in 1955 1957<br />

(Marshall, 1960).<br />

Control of the loose cereal smuts was impossible<br />

until the Danish plant pathologist Jens<br />

Ludwig Jensen invented the hot water treatment<br />

method in the 1880s (Large, 1940; Ainsworth,<br />

1981). This was based on the observation that<br />

Ustilago spp. are less heat-<strong>to</strong>lerant than cereals,<br />

and infected seeds could be effectively disinfected<br />

by soaking them first in cold water for<br />

5 h followed by a dip in hot water; 10 min at<br />

54°C for wheat and 15 min at 52°C for barley<br />

(Fischer & Hol<strong>to</strong>n, 1957; Ainsworth, 1981).<br />

Today, effective fungicidal seed dressings are<br />

available and these usually combine systemic<br />

fungicides such as carboxin (Fig. 23.13) or its<br />

derivatives with protectant fungicides such<br />

as captan, maneb or pentachloronitrobenzene<br />

(Kulka & von Schmeling, 1995). Such seed<br />

dressings control loose and covered smuts, in<br />

addition <strong>to</strong> numerous other fungal diseases.<br />

Since infection of next season’s grain with<br />

loose smuts occurs at flowering, one obvious<br />

method of control is <strong>to</strong> inspect crops grown for<br />

seed at flowering time and <strong>to</strong> assess the<br />

incidence of smutted heads. In these so-called<br />

seed certification schemes, only crops which<br />

contain fewer than a defined limit, e.g. one<br />

smutted ear in 10 000 ears, are approved for use<br />

as seed s<strong>to</strong>cks (Doling, 1966). It is also possible <strong>to</strong><br />

detect the presence of loose smut mycelium<br />

within the embryos by microscopic examination<br />

(Mor<strong>to</strong>n, 1961) or PCR-based methods (Pearce,<br />

1998). The latter can also detect the spores of any<br />

other fungus of interest on the seed coat,<br />

including covered smuts such as T. caries,<br />

T. controversa and T. indica. PCR-based methods<br />

are being developed rapidly, partly because<br />

of the need for rapid testing of exported or<br />

imported agricultural produce for contamination,<br />

and partly for the purpose of thwarting<br />

bioterrorist attacks (Schaad et al., 2003).<br />

If all else fails, many systemic fungicides,<br />

e.g. sterol demethylation inhibi<strong>to</strong>rs (see p. 410),<br />

are effective against the systemic smuts when<br />

sprayed on<strong>to</strong> the growing crop. In practice, these<br />

fungicides are often applied <strong>to</strong> control infections<br />

caused by non-smut cereal pathogens, with the<br />

suppression of Ustilago spp. as an additional<br />

bonus which often goes unnoticed by the<br />

farmer (Jones, 1999).<br />

Although many crop plants and their wild<br />

relatives possess resistance genes against smut<br />

fungi, smut-resistant cultivars are less important<br />

in an agricultural context than those with<br />

resistance against other biotrophic pathogens,<br />

e.g. rusts (see pp. 625 and 627) or powdery<br />

Fig 23.13 Molecular structure of carboxin, a systemic<br />

fungicide commonly applied as a seed dressing.Carboxin acts<br />

by inhibiting the enzyme succinate dehydrogenase (complex II)<br />

in fungal mi<strong>to</strong>chondrial respiration (see Uesugi,1998).


652 USTILAGINOMYCETES: SMUT FUNGI AND THEIR ALLIES<br />

mildews (p. 408). Avirulence genes involved in<br />

gene-for-gene interactions have been desribed<br />

for several cereal smuts (Sidhu & Person, 1972;<br />

Eckstein et al., 2002; Hu et al., 2003). The<br />

hypersensitive response in incompatible interactions<br />

is often microscopically small (Hu et al.,<br />

2003). Field resistance against smut fungi may<br />

also occur as a combination of several genes<br />

(Nelson et al., 1998). An interesting case of singlegene<br />

resistance has been reported by Wilson<br />

and Hanna (1998) as trichome-less mutants of<br />

pearl millet (Pennisetum glaucum) showing a 50%<br />

reduction in disease severity caused by the<br />

smut Moesziomyces penicillariae. Curiously little is<br />

known about resistance mechanisms in Ustilago<br />

maydis despite its status as a well-examined<br />

‘model organism’ and the availability of its<br />

genome sequence.<br />

23.3 Microbotryales<br />

(Urediniomycetes)<br />

The order Microbotryales contains smut fungi<br />

with features almost indistinguishable from<br />

those described in the preceding section. This<br />

includes their general life cycle of teliospores<br />

germinating by means of a promycelium which<br />

produces haploid yeast-like sporidia, and the<br />

infection of host plants following establishment<br />

of a dikaryotic infection hypha. Intriguingly,<br />

however, phylogenetic studies have clearly<br />

shown the Microbotryales <strong>to</strong> belong <strong>to</strong> the<br />

Urediniomycetes (Blanz & Gottschalk, 1984;<br />

Begerow et al., 1997; Swann et al., 1999). There<br />

are certain microscopic details which distinguish<br />

them from Ustilaginomycetes, notably that<br />

growth by Microbotryum and allied species<br />

in planta is exclusively intercellular (Bauer et al.,<br />

1997) and that the teliospores have a violetpurplish<br />

rather than brown pigmentation.<br />

The most important genus in the Microbotryales<br />

is Microbotryum which currently<br />

contains about 75 species (Vánky, 1998, 1999).<br />

All of them parasitize dicotyledonous hosts, and<br />

they can be resolved by DNA sequence analysis<br />

in<strong>to</strong> two groups (Almaraz et al., 2002). The original<br />

genus Microbotryum parasitizes the anthers of<br />

plants belonging <strong>to</strong> the Caryophyllaceae, whereas<br />

species of the second group cause smuts on<br />

various organs of other host plants. Until<br />

recently, members of the second group were<br />

considered <strong>to</strong> belong <strong>to</strong> Ustilago. The eventful<br />

taxonomic his<strong>to</strong>ry of these ‘dicot Ustilago’ spp. has<br />

been recounted by Vánky (1998, 1999); based on<br />

the findings of Almaraz et al. (2002) they will<br />

probably have <strong>to</strong> be given a new generic name,<br />

adding at least one further twist <strong>to</strong> the s<strong>to</strong>ry.<br />

Good descriptions of many ‘dicot Ustilago’ spp.<br />

have been given by Vánky (1994, 1998). We shall<br />

devote our attention <strong>to</strong> Microbotryum sensu stric<strong>to</strong>,<br />

cause of anther smut on Caryophyllaceae.<br />

23.3.1 Microbotryum violaceum<br />

Anther smut of Caryophyllaceae has long<br />

fascinated biologists with a wide range of<br />

research interests. Although infection is systemic<br />

throughout the host plant, the disease symp<strong>to</strong>ms<br />

are confined <strong>to</strong> the anthers in which pollen is<br />

replaced by purple teliospores. Several members<br />

of the Caryophyllaceae are infected, and there is<br />

evidence that Microbotryum violaceum (formerly<br />

called Ustilago violacea), the causal fungus, is<br />

undergoing speciation in diverse host plants<br />

(An<strong>to</strong>novics et al., 2002), so that species delimitation<br />

is difficult at present (see Vánky, 1994). In<br />

dioecious hosts such as red or white campion<br />

(Silene dioica and S. alba), the flowers of infected<br />

female plants are stimulated <strong>to</strong> produce anthers,<br />

i.e. their morphology changes from female <strong>to</strong><br />

male. A thorough review of the general biology<br />

of M. violaceum has been written by Day and<br />

Garber (1988).<br />

Life cycle and genetic recombination<br />

The diploid teliospores of M. violaceum are<br />

transported from diseased <strong>to</strong> healthy flowers<br />

by pollinating insects. They germinate by forming<br />

a promycelium (Fig. 23.14) which often<br />

separates in<strong>to</strong> an exterior three-celled section<br />

and a one-celled fragment remaining in the<br />

teliospore. Each promycelium segment produces<br />

sporidia. The three-celled segment becomes<br />

readily detached from the teliospore and may<br />

continue <strong>to</strong> develop sporidia after separation.


MICROBOTRYALES (UREDINIOMYCETES)<br />

653<br />

Fig 23.14 Microbotryum violaceum.<br />

(a) Germinating teliospore showing three cells<br />

of the promycelium. (b) Detached three-celled<br />

promycelium segment producing sporidia.<br />

(c) Detached sporidia reproducing as yeast<br />

cells.<br />

In nutrient-rich conditions and at temperatures<br />

above 25°C, the haploid sporidia reproduce by<br />

yeast-like budding, whereas mating occurs at<br />

lower temperatures in nutrient-poor media<br />

such as tapwater agar. Microbotryum violaceum is<br />

heterothallic and bipolar, and it was the first<br />

smut fungus for which such an incompatibility<br />

system was demonstrated (Kniep, 1919). During<br />

mating, an a2 sporidium or yeast cell produces a<br />

conjugation tube which grows <strong>to</strong>wards the a1<br />

cell. Provided that a-<strong>to</strong>copherol (vitamin E) is<br />

present, fusion is followed by the production of a<br />

dikaryotic hypha which infects the flower of the<br />

host plant (Castle & Day, 1984).<br />

In the absence of a-<strong>to</strong>copherol or on nutrientrich<br />

media, de-dikaryotization occurs and<br />

haploid or aneuploid yeast cells are formed by<br />

the loss of chromosomes. Several other complications<br />

exist in the life cycle of M. violaceum,<br />

illustrating the wealth of non-meiotic mechanisms<br />

of recombination encountered in fungi<br />

(Day & Garber, 1988; Day, 1998). For instance,<br />

nuclear fusion not associated with meiosis is<br />

readily inducible in M. violaceum, and this<br />

generates somatic diploids. Mi<strong>to</strong>tic crossing-over<br />

occurs at a high frequency in somatic diploids.<br />

As already mentioned, M. violaceum has<br />

dimorphic mating type chromosomes (Hood,<br />

2002), and whilst the mating type locus is a<br />

hot spot for mi<strong>to</strong>tic recombination, meiotic<br />

recombination does not occur in this region of<br />

the genome.<br />

Somatic diploids can be mated if they are<br />

homozygous for mating type, so that triploid or<br />

tetraploid strains can be generated. In nutrientrich<br />

media, these strains grow as yeast cells but<br />

are unstable due <strong>to</strong> the gradual loss of chromosomes.<br />

The size of the yeast cells is correlated<br />

with their ploidy level. There is also evidence for<br />

the activity of transposons in enhancing genetic<br />

recombination in M. violaceum (Garber & Ruddat,<br />

2002). The genetic flexibility of M. violaceum may<br />

explain the existence of numerous strains in<br />

nature.<br />

Fimbriae and conjugation<br />

The process of sporidial conjugation has been<br />

extensively studied in M. violaceum. Within 2 h of<br />

placing sporidia of opposite mating type on a<br />

suitable medium, a series of events occurs which<br />

culminates in the formation of conjugation tubes<br />

and plasmogamy. Poon and Day (1974) observed<br />

an early adhesion between the two sporidia<br />

before any wall-<strong>to</strong>-wall contact was established,<br />

and they discovered that this initial contact<br />

was mediated by fimbriae (Fig. 23.15) whose<br />

existence in fungi had not been appreciated<br />

previously. Subsequent investigations revealed<br />

that fungi from most major groups (Zygomycota,<br />

Ascomycota, Basidiomycota) possess such


654 USTILAGINOMYCETES: SMUT FUNGI AND THEIR ALLIES<br />

Fig 23.15 Fimbriae of M. violaceum. Shadowing-TEM<br />

view. Reprinted from Gardiner and Day (1988) by<br />

copyright permission of the National Research<br />

Council of Canada; original print kindly provided by<br />

A.W. Day.<br />

fimbriae, and that these have a similar biochemical<br />

composition. Fimbriae can be of variable<br />

length (up <strong>to</strong> 20 mm) but have a uniform diameter<br />

of 7 nm. Most of the biochemical work on fungal<br />

fimbriae was done with M. violaceum, and this<br />

subject has been well reviewed by Celerin and Day<br />

(1998).<br />

Most fungal fimbriae consist of three structural<br />

components. The main component is<br />

proteinaceous, with strong homology <strong>to</strong> collagen<br />

which had been thought previously <strong>to</strong> exist only<br />

in animals (Celerin et al., 1996). The second<br />

component consists of polysaccharide glycosylation<br />

chains on the collagen protein. Whereas<br />

deglycosylated fungal collagen can still polymerize<br />

<strong>to</strong> form fimbriae, these become unstable in<br />

extreme conditions. Further, the glycosylation<br />

chains are involved in the specificity of mating<br />

recognition in M. violaceum, with the fimbrial<br />

protein component of a1 sporidia binding <strong>to</strong><br />

the mannose residues on the proteins of the<br />

a2 fimbriae, i.e. acting as a lectin (Castle et al.,<br />

1996). The third component of fungal fimbriae is<br />

a short (30 nucleotides) single-stranded fimbrial<br />

RNA (fRNA), the function of which is as yet<br />

unknown (Celerin et al., 1994). The fimbriae of<br />

M. violaceum do not traverse the cell wall, but<br />

are anchored <strong>to</strong> it by means of a protein which<br />

may be functionally analogous <strong>to</strong> mammalian<br />

fibrinogen.<br />

Whereas M. violaceum produces only one<br />

type of fimbria, other fungi have several<br />

morphologically distinguishable types (Gardiner<br />

& Day, 1988). In Candida albicans (p. 277), fimbriae<br />

play an important part in adhesion <strong>to</strong> human<br />

tissue and initiation of infection, and the<br />

fimbriae involved in this process are not made<br />

up of collagen (Yu et al., 1994).<br />

Host pathogen interactions<br />

Initial infection of the host plants by M. violaceum<br />

occurs if a pollinating insect carries teliospores<br />

from a diseased <strong>to</strong> a healthy flower (Jennersten,<br />

1988). Anther smut is thus a sexually transmitted<br />

disease. The entire flower is colonized by intercellular<br />

dikaryotic mycelium. Infection results<br />

in reproductive sterility of the host; in male<br />

flowers, the anthers become converted <strong>to</strong> teliospore<br />

sori whereas in female flowers, the female<br />

traits are suppressed and anthers are formed<br />

which carry teliospores. This development has<br />

been described in detail for infections of Silene<br />

latifolia (Uchida et al., 2003). Diploid female<br />

plants carry two X chromosomes in addition <strong>to</strong><br />

22 au<strong>to</strong>somes, whereas male plants contain<br />

22 au<strong>to</strong>somes plus one X and one Y chromosome.<br />

The Y chromosome encodes genes that suppress<br />

the development of female traits and promote<br />

stamen development, thereby making the flower<br />

male. Microbotryum violaceum must therefore produce<br />

a signal called male-sterility res<strong>to</strong>ration<br />

fac<strong>to</strong>r that substitutes for the Y chromosome in<br />

female flowers (Uchida et al., 2003). Its identity is<br />

as yet unknown, but it must be a signal molecule


EXOBASIDIALES (USTILAGINOMYCETES)<br />

655<br />

or transcription fac<strong>to</strong>r acting on genes encoded<br />

on the au<strong>to</strong>somes or the X chromosome, since<br />

the expression patterns of some male-specific<br />

genes in healthy male and infected female<br />

flowers are similar (Scutt et al., 1997).<br />

In perennial host species, M. violaceum can<br />

colonize the entire host in the course of<br />

1 2 years, overwintering systemically in the<br />

root system (Alexander & An<strong>to</strong>novics, 1988).<br />

Infected plants acquire lifelong sterility, i.e.<br />

they will continue <strong>to</strong> produce flowers but these<br />

will all be male, developing anthers in which<br />

pollen grains are replaced by teliospores.<br />

The ecology of M. violaceum<br />

The ecological effects of M. violaceum infections<br />

have been considered by several workers.<br />

Systemically infected host plants flower earlier<br />

than healthy plants, and this may force pollina<strong>to</strong>rs<br />

such as bumblebees <strong>to</strong> visit infected flowers<br />

early in the season. Consequently, teliospore<br />

dispersal peaks earlier than pollen dispersal<br />

(Jennersten, 1988). In direct comparison, diseased<br />

flowers are less attractive <strong>to</strong> bumblebees<br />

(Shykoff & Bucheli, 1995), possibly because they<br />

contain less nectar and are asymmetric and<br />

smaller than healthy flowers (Shykoff & Kaltz,<br />

1998). A pollina<strong>to</strong>r preference for healthy flowers<br />

may be advantageous <strong>to</strong> the pathogen because it<br />

enhances the opportunity for spreading the<br />

infection following an accidental visit <strong>to</strong> a<br />

diseased plant (Shykoff & Bucheli, 1995).<br />

Although systemically infected host plants<br />

produce an increased number of flowers which<br />

also remain open for longer, the drawback is<br />

that the root biomass is decreased, thereby<br />

potentially affecting the chance of the pathogen<br />

<strong>to</strong> overwinter systemically. Since healthy<br />

female flowers remain open for much longer<br />

than healthy male flowers, they are at greater<br />

risk of infection (Kaltz & Shykoff, 2001). As in<br />

many venereal diseases, a more attractive display<br />

draws a greater number of visi<strong>to</strong>rs, and one<br />

possible adaptation <strong>to</strong> M. violaceum is predicted <strong>to</strong><br />

be a reduction in flower size. Another might be<br />

the change from a perennial <strong>to</strong> an annual habit,<br />

as this would prevent overwintering of the<br />

pathogen (Kaltz & Shykoff, 2001). In situations<br />

where perennial hosts are forced in<strong>to</strong> an annual<br />

habit, e.g. near field borders or railway lines<br />

subject <strong>to</strong> regular clearing, infections by<br />

M. violaceum are rare or absent (Alexander &<br />

An<strong>to</strong>novics, 1988). The interaction between<br />

M. violaceum and its hosts may therefore have<br />

different outcomes, ranging from local extinction<br />

of the host <strong>to</strong> that of the pathogen.<br />

23.4 Exobasidiales<br />

(Ustilaginomycetes)<br />

The order Exobasidiales (subclass Exobasidiomycetidae)<br />

is a well-defined group (Begerow et al.,<br />

1997, 2002) of ecologically obligate, biotrophic<br />

plant pathogens containing less than 100 species.<br />

Most species cause systemic or limited infections<br />

of shoots or leaves, and these are often accompanied<br />

by hypertrophy of the infected tissue.<br />

The most important genus is Exobasidium. <strong>Fungi</strong><br />

grouped in the Exobasidiales firmly belong <strong>to</strong><br />

the Ustilaginomycetes and possess a haploid<br />

saprotrophic yeast-like phase and a dikaryotic<br />

biotrophic phase. However, they differ from the<br />

‘true smuts’ (Ustilaginomycetidae) in several<br />

aspects:<br />

1. They do not produce teliospores, but<br />

basidia are formed directly on the surface of<br />

the infected host.<br />

2. The basidium appears similar <strong>to</strong> that of<br />

the Homobasidiomycetes, i.e. it looks like a<br />

holobasidium, not divided in<strong>to</strong> a pro- and metabasidium<br />

as in smut fungi (but see below).<br />

3. Basidiospores are violently discharged<br />

using the surface-tension catapult mechanism,<br />

and they often become septate after discharge.<br />

4. The mycelium is usually intercellular,<br />

with haus<strong>to</strong>ria often present. These differ from<br />

the intracellular hyphae of smut fungi in not<br />

being completely surrounded by a sheath.<br />

Instead, a complex interaction apparatus is<br />

formed at the haus<strong>to</strong>rial apex (Fig. 23.17).<br />

23.4.1 Exobasidium<br />

About 50 Exobasidium spp. are known. All of<br />

them cause either local infections of individual<br />

leaves or more widespread, sometimes systemic<br />

infections of whole shoots or shoot tips


656 USTILAGINOMYCETES: SMUT FUNGI AND THEIR ALLIES<br />

Fig 23.16 Exobasidium japonicum. (a) Basidium on the lower surface of an evergreen azalea. (b) Freshly discharged aseptate<br />

basidiospores. (c) Basidiospore after a few hours’ incubation on pota<strong>to</strong> dextrose agar.The spore has formed a transverse septum<br />

and has germinated. (d) Basidiospore after12 h.The germ tubes have branched and are budding off elongated yeast cells.<br />

Fig 23.17 Lobed haus<strong>to</strong>rium of Exobasidium sp.Three<br />

interaction sites with the host cy<strong>to</strong>plasm are numbered. Site1<br />

is shown in median section, and the interaction ring is visible<br />

(arrowheads).Sites2and3areintangentialsection,showing<br />

electron-dense deposits (d).Tubulovesicular cy<strong>to</strong>plasmic<br />

structures involved in secretion of the electron-dense matter<br />

are also visible (arrows).The penetration site of the host wall<br />

is at the bot<strong>to</strong>m of the picture (double arrowheads).<br />

Reprinted from Begerow et al.(2002),Mycological Progress,<br />

with permission of IHW-Verlag; original print kindly provided<br />

by D. Begerow.<br />

(Plates 12i,j). Infection commonly gives rise <strong>to</strong><br />

hypertrophied tissue. Keys and descriptions have<br />

been given by Nannfeldt (1981) and Ing (1998).<br />

Frequently infected host plants include members<br />

of the Ericaceae, e.g. Vaccinium (bilberry, blueberry)<br />

in heaths, moors and forests, and ornamental<br />

azaleas and Rhododendron spp. in gardens.<br />

Cosmopolitan examples are E. japonicum on<br />

evergreen azaleas (Plate 12i; Fig. 23.16) and<br />

E. vaccinii on Vaccinium spp. (Plate 12j). A second<br />

group of host plants is the Theaceae family, with<br />

the tea plant (Camellia sinensis) a prominent<br />

casualty of E. vexans which adversely affects the<br />

quantity and quality of the tea harvest (Gulati<br />

et al., 1999).<br />

Infections by Exobasidium spp. are easily<br />

confused with symp<strong>to</strong>ms caused by Taphrina<br />

(Plate 4a,c), and these two genera, although<br />

phylogenetically unrelated, have much in<br />

common. Plant tissues infected by Exobasidium<br />

are often swollen and show a pale or reddish<br />

discoloration. There is an obvious hormonal<br />

imbalance in the infected host tissue, although<br />

no detailed examinations have been carried out.<br />

Infected shoots are often affected in their reproductive<br />

development (Wolfe & Rissler, 1999).<br />

A whitish superficial layer of basidia is eventually<br />

produced on the surface of infected host<br />

tissue, and these carry about 2 8 sterigmata


EXOBASIDIALES (USTILAGINOMYCETES)<br />

657<br />

(Fig. 23.16a). Basidiospores are initially aseptate<br />

(Fig. 23.16b), but in many species they become<br />

septate after their discharge, and both cells can<br />

germinate by emitting germ tubes (Fig. 23.16c).<br />

Elongated haploid yeast cells are produced from<br />

these germ tubes or directly from the basidiospore<br />

cell (Fig. 23.16d). Several Exobasidium spp.<br />

grow as yeast cells in culture. Infection of host<br />

plants may occur from basidiospores or yeast<br />

cells and gives rise <strong>to</strong> a dikaryotic intercellular<br />

mycelium (Mims & Nickerson, 1986). Exobasidium<br />

spp. may overwinter systemically in their host,<br />

or as spores on the outside, e.g. in bud scales<br />

or bark.<br />

Haus<strong>to</strong>ria are formed by the intercellular<br />

hyphae, and these show unique characteristics<br />

in being lobed and producing an electron-dense<br />

apical ring at the localized point of contact<br />

with the host plasmalemma. This then becomes<br />

elaborated in<strong>to</strong> an apical cap which is associated<br />

with the host wall (Bauer et al., 1997) or<br />

directly with the host plasmalemma (Mims,<br />

1982). The ring and cap material is secreted<br />

by the haus<strong>to</strong>rium from an elaborate tubular<br />

membrane system (Fig. 23.17). This cap may be<br />

homologous <strong>to</strong> the thick sheath which surrounds<br />

the intracellular hyphae of smut fungi (see<br />

Fig. 23.6).<br />

Although the basidia of Exobasidium look<br />

like typical club-shaped holobasidia, their<br />

development differs from that in the homobasidiomycetes<br />

(Mims et al., 1987). Nuclear fusion<br />

occurs in a subterminal cell (strictly speaking<br />

the probasidium), and then the nucleus migrates<br />

in<strong>to</strong> the basidium proper where meiosis is<br />

completed. The basidium of Exobasidium is thus<br />

a metabasidium in disguise, and the pattern of<br />

basidial development is equivalent <strong>to</strong> that in the<br />

rust and smut fungi. A further unusual feature<br />

is that the hilar appendices of basidiospores on<br />

the basidium point outwards (Fig. 23.16a), not<br />

inwards as in Homobasidiomycetes.


24<br />

Basidiomycete yeasts<br />

24.1 <strong>Introduction</strong><br />

<strong>Fungi</strong> living predominantly or exclusively as<br />

yeasts are encountered in three classes of<br />

Basidiomycota, namely the Heterobasidiomycetes<br />

(Chapter 21), Urediniomycetes (Chapter 22)<br />

and Ustilaginomycetes (Chapter 23). We shall<br />

discuss basidiomycete yeasts <strong>to</strong>gether in the<br />

present chapter because these organisms,<br />

although taxonomically diverse, are unified by<br />

many biological features.<br />

24.1.1 Ecology<br />

Little is known about the ecology of basidiomycete<br />

yeasts. They occur in marine and freshwater<br />

habitats, the soil and the plant rhizosphere, and<br />

especially on above-ground plant surfaces such<br />

as tree bark, leaves, flowers and fruits. A certain<br />

degree of specificity of yeast species relative <strong>to</strong><br />

plant species or organs of a given plant host has<br />

been observed (Phaff, 1990). Basidiomycete yeasts<br />

may be found in all climatic zones from the<br />

arctic <strong>to</strong> the tropics. They generally exist as<br />

saprotrophic phylloplane organisms. When<br />

nutrients become available, there may be a<br />

steep increase in the population density of<br />

these yeasts. Many yeasts isolated from soil<br />

have their origin in vegetation which becomes<br />

incorporated in<strong>to</strong> humus after leaf fall.<br />

Basidiomycete yeasts are not noted as plant<br />

pathogens, but some species can infect animals.<br />

Cryp<strong>to</strong>coccus neoformans is one of the most serious<br />

fungal pathogens of humans (pp. 661 665).<br />

Members of another group (Malassezia spp.) live<br />

commensally on the skin of humans and other<br />

mammals, causing superficial derma<strong>to</strong>mycoses<br />

under suitable conditions (p. 671).<br />

Basidiomycete yeasts therefore share many<br />

ecological features with their ascomycete counterparts,<br />

and they are easily isolated following<br />

similar procedures, i.e. by plating out soil<br />

suspensions, leaf washings or filters bearing<br />

water samples on<strong>to</strong> standard agar media<br />

augmented with antibacterial antibiotics<br />

(p. 262). One good way <strong>to</strong> isolate ballis<strong>to</strong>conidium-forming<br />

yeasts is <strong>to</strong> attach a piece of<br />

vegetation <strong>to</strong> the underside of a Petri dish lid,<br />

permitting ballis<strong>to</strong>sporic yeasts <strong>to</strong> shower their<br />

spores on<strong>to</strong> the agar medium. Within 2 3 days,<br />

yeasts such as Sporobolomyces spp. will grow and<br />

can be isolated in pure culture. Yeasts are<br />

preserved in an active state on agar slopes<br />

at 4°C. Lyophilized preparations can also be<br />

made from vegetative cells of many species,<br />

and these remain viable for several years.<br />

Many basidiomycete yeasts, the so-called<br />

‘red yeasts’, are coloured yellow, orange, pink<br />

or red due <strong>to</strong> the presence of carotenoids (see<br />

Fig. 24.8). Red yeasts are found in all three classes<br />

containing basidiomycete yeasts, but they are<br />

rare among ascomycetes (see p. 253 for the<br />

Lalaria state of Taphrina). Carotenoid production<br />

can be of commercial value, as in Phaffia<br />

rhodozyma which produces astaxanthin, an<br />

important food and feed pigment (p. 665).<br />

Other commercial applications of basidiomycete<br />

yeasts are as potential producers of lipid, e.g. as<br />

cocoa butter substitute (Ratledge, 1997), and as<br />

biocontrol agents against s<strong>to</strong>rage rots of fruits


INTRODUCTION<br />

659<br />

Fig 24.1 Basidiomycete yeasts<br />

as seen by light microscopy.<br />

(a) Budding in Dioszegia sp.<br />

(Heterobasidiomycetes).The yeast<br />

cells are isodiametric with a slightly<br />

pointed bud region. (b) Budding<br />

cells of Rhodo<strong>to</strong>rulaglutinis<br />

(Urediniomycetes).This species<br />

produces a prominent polysaccharide<br />

capsule which has been revealed<br />

by mounting yeast cells in Indian<br />

ink. (c) Collarette at the site of<br />

repeated bud formation in<br />

Sakaguchia dacryoidea<br />

(Urediniomycetes). (d) Annellidic<br />

structure formed by a member of<br />

the Dioszegia hungarica group<br />

(Heterobasidiomycetes).<br />

(e) Ballis<strong>to</strong>conidium formation<br />

in Sporobolomycesroseus<br />

(Urediniomycetes). (f) Septate<br />

stalked basidium arising from a<br />

teliospore of Sporidiobolus ruineniae<br />

(Urediniomycetes). (a e) <strong>to</strong> same<br />

scale.Original print of f kindly<br />

provided by J. W. Fell.<br />

caused by filamen<strong>to</strong>us fungi (Janisiewicz &<br />

Korsten, 2002).<br />

Detailed descriptions of basidiomycete yeasts<br />

along with their ascomycete counterparts may<br />

be found in Kurtzman and Fell (1998) and<br />

Barnett et al. (2000). A useful general introduction<br />

<strong>to</strong> the <strong>to</strong>pic is that by Fell et al. (2001).<br />

24.1.2 Morphology and life cycles<br />

Asexual reproduction is mainly by budding. In<br />

contrast <strong>to</strong> ascomycete yeasts, the budding sites<br />

are confined <strong>to</strong> either or both poles of the<br />

vegetative cells which are usually ellipsoid or<br />

elongated (Figs. 24.1b,d) but may be isodiametric<br />

(Fig. 24.1a). Yeast cells may be coated by a<br />

prominent polysaccharide capsule (Fig. 24.1b),<br />

giving the culture a slimy appearance. Repeated<br />

budding at the same site may lead <strong>to</strong> the<br />

formation of collarettes (Fig. 24.1c) or even<br />

annellide-like structures (Fig. 24.1d). The formation<br />

of daughter cells is enteroblastic, i.e. only<br />

the inner wall of the mother cell extends <strong>to</strong> form<br />

the wall of the daughter cell. The whole budding<br />

yeast cell can thus be viewed as a phialide. This is<br />

in contrast <strong>to</strong> most ascomycete yeasts which bud<br />

in a holoblastic manner (i.e. the daughter wall is<br />

continuous with the entire wall of the mother<br />

cell). Given a little practice, it is often possible<br />

with the light microscope <strong>to</strong> recognize a basidiomycete<br />

yeast as such. Transmission electron<br />

microscopy has revealed differences between<br />

the two-layered cell walls of ascomycetes and<br />

the multi-layered lamellate walls of basidiomycete<br />

yeasts (Kreger-van Rij & Veenhuis, 1971).<br />

These differences have been correlated with the<br />

reaction of basidiomycete but not ascomycete<br />

yeasts with the diazonium blue B stain (Simmons<br />

& Ahern, 1987).<br />

Apart from budding, several genera of basidiomycete<br />

yeasts such as Sporobolomyces and its<br />

teleomorph Sporidiobolus (Urediniomycetes) or<br />

Bullera and its teleomorph Bulleromyces


660 BASIDIOMYCETE YEASTS<br />

(Heterobasidiomycetes) also produce conidia<br />

which are actively liberated in the manner of<br />

basidiospores, i.e. by means of the surfacetension<br />

catapult mechanism involving Buller’s<br />

drop (Fig. 24.1e). Since these are asexual propagules,<br />

they are called ballis<strong>to</strong>conidia. Their<br />

existence provided one of the first clues that<br />

the yeasts producing them belong <strong>to</strong> the Basidiomycota<br />

(Kluyver & van Niel, 1927).<br />

Sexual reproduction is relatively rare in<br />

basidiomycete yeasts except, of course, for those<br />

species which have dominant mycelial diploid or<br />

dikaryotic stages such as the smut fungi<br />

discussed in Chapter 23, or jelly fungi and allies<br />

(Chapter 21). Following mating between compatible<br />

yeast cells, a limited dikaryotic mycelium<br />

often bearing clamp connections may arise, and<br />

this produces basidia either directly or, more<br />

commonly, via thick-walled resting cells called<br />

teliospores. There, karyogamy occurs, i.e. the<br />

teliospores function as probasidia. Teliospores<br />

germinate in a manner described in detail for<br />

rust and smut fungi, namely by the production<br />

of a promycelium (¼ metabasidium) which<br />

may or may not undergo transverse septation.<br />

The basidiospores thus produced germinate by<br />

budding as yeast cells (Fig. 24.1f).<br />

Where present, septa can be examined by<br />

transmission electron microscopy for their ultrastructure,<br />

and this is an important feature in<br />

classification. In general, the septa of yeasts<br />

belonging <strong>to</strong> Urediniomycetes and Ustilaginomycetes<br />

contain simple pores whereas those of<br />

heterobasidiomycete yeasts have dolipores, often<br />

with a parenthesome (Fell et al., 2001).<br />

24.1.3 Phylogeny of basidiomycete yeasts<br />

In addition <strong>to</strong> examining morphological features<br />

as outlined above, several biochemical tests can<br />

be performed <strong>to</strong> characterize yeasts. Such tests,<br />

e.g. carbon source utilization, the identity of<br />

coenzyme Q or the spectrum of killer <strong>to</strong>xins<br />

produced, have been employed extensively in<br />

the past <strong>to</strong> identify yeasts and are still relevant<br />

<strong>to</strong>day (Yarrow, 1998; Fell et al., 2001). However, the<br />

results are rarely clear-cut, and taxonomic confusion<br />

has resulted due <strong>to</strong> the extensive overlap<br />

of features between members of different taxa.<br />

Hence, the names of many basidiomycete yeasts<br />

are of descriptive rather than taxonomic value<br />

and may be found in several phylogenetically<br />

distinct clades (Table 24.1). For instance, the<br />

main feature <strong>to</strong> distinguish Sporobolomyces from<br />

Rhodo<strong>to</strong>rula is the presence or absence (respectively)<br />

of ballis<strong>to</strong>conidia, and it is now known<br />

that these character states, and hence the two<br />

generic names, are of little taxonomic relevance.<br />

Major work is currently being carried out in<br />

order <strong>to</strong> establish phylogenetically coherent<br />

groups of yeasts, using especially ribosomal<br />

DNA but also increasingly other gene sequences.<br />

Valuable recent contributions <strong>to</strong> phylogeny at<br />

higher taxonomic levels are those by Fell et al.<br />

(2000, 2001) and Scorzetti et al. (2002) in which<br />

the heterobasidiomycete, urediniomycete and<br />

ustilaginomycete clades of yeasts have been<br />

circumscribed (Table 24.1). The integration of<br />

some of these clades in<strong>to</strong> the taxonomy of<br />

filamen<strong>to</strong>us basidiomycetes has yet <strong>to</strong> be accomplished.<br />

Additionally, numerous publications<br />

have dealt with the analysis of taxa at the level<br />

of genus or species. Hence, although it is<br />

generally estimated that only 1 5% of all<br />

basidiomycete yeasts have been discovered as<br />

yet, a large database of DNA sequences is already<br />

available. A convenient side effect of this work<br />

is that the identification of new isolates or at<br />

least their assignment <strong>to</strong> a given family or genus<br />

is a relatively straightforward matter if their<br />

rDNA sequences can be obtained. Biochemical<br />

tests and microscopy can then be used <strong>to</strong> verify<br />

and extend the identification.<br />

24.2 Heterobasidiomycete yeasts<br />

This is a large and morphologically diverse<br />

group of yeasts. We shall focus on two species<br />

which have been particularly well examined, the<br />

serious human pathogen Filobasidiella neoformans<br />

and the astaxanthin producer Phaffia rhodozyma.<br />

Trichosporon spp., which are occasional opportunistic<br />

human pathogens, are not further<br />

discussed. They show a diversity of growth<br />

forms, including hyphae, pseudohyphae, yeast<br />

cells, blas<strong>to</strong>conidia and arthroconidia.


HETEROBASIDIOMYCETE YEASTS<br />

661<br />

Table 24.1. Some important anamorph (A) and<br />

teleomorph (T) genera of basidiomycete yeasts and<br />

their approximate taxonomic position according<br />

<strong>to</strong> Fell et al. (2000, 2001) and Scorzetti et al. (2002).<br />

Heterobasidiomycetes (Section 24.2)<br />

1. Tremellales<br />

Fellomyces A<br />

Bulleromyces T ¼ Bullera A<br />

Cryp<strong>to</strong>coccus A<br />

Dioszegia A<br />

Filobasidiella neoformans T ¼Cryp<strong>to</strong>coccus<br />

neoformans A(p.661)<br />

Tremellayeast states (p. 605)<br />

2. Trichosporonales (exclusively anamorphic fungi)<br />

Cryp<strong>to</strong>coccus A<br />

Trichosporon A<br />

3. Filobasidiales<br />

Filobasidium T ¼Cryp<strong>to</strong>coccus A<br />

4. Cys<strong>to</strong>filobasidiales<br />

Cys<strong>to</strong>filobasidium T ¼Cryp<strong>to</strong>coccus A<br />

Itersonilia perplexans A(seep.493)<br />

Xanthophyllomyces dendrorhous Tand<br />

Phaffia rhodozyma A(p.665)<br />

Urediniomycetes (Section 24.3)<br />

1. Microbotryum clade<br />

Microbotryum yeast states (see p. 652)<br />

Rhodo<strong>to</strong>rula A<br />

2. Sporidiobolus clade<br />

Rhodosporidium T ¼ Rhodo<strong>to</strong>rula A<br />

Sporidiobolus T ¼ Sporobolomyces A (p.666)<br />

3. Erythrobasidium clade<br />

Erythrobasidium T<br />

Sakaguchia T<br />

Rhodo<strong>to</strong>rula A<br />

Sporobolomyces A<br />

4. Agaricostilbum clade<br />

Agaricostilbum T<br />

Bensing<strong>to</strong>nia A<br />

Sporobolomyces A<br />

Ustilaginomycetes (Section 24.4)<br />

1. Ustilaginomycetidae<br />

Pseudozyma A<br />

Ustilago yeast states (see p. 636)<br />

2. Exobasidiomycetidae<br />

Exobasidiumyeast states (see p.655)<br />

Rhodo<strong>to</strong>rula A<br />

Tilletia yeast states (see p.636)<br />

Tilletiopsis A (yeast states of smutfungi)<br />

3. Malasseziales<br />

Malassezia A(p.670)<br />

24.2.1 Filobasidiella (Cryp<strong>to</strong>coccus)<br />

neoformans<br />

The anamorph genus Cryp<strong>to</strong>coccus comprises some<br />

30 40 species which are scattered throughout<br />

all 4 yeast-containing orders of the Heterobasidiomycetes,<br />

and with a range of different teleomorphs<br />

(Table 24.1). Clearly, therefore, the genus<br />

is polyphyletic. Cryp<strong>to</strong>coccus spp. are ubiqui<strong>to</strong>us,<br />

being found in all climatic zones on plant<br />

material and in the soil. One species, C. neoformans<br />

(teleomorph Filobasidiella neoformans), is a human<br />

pathogen and among the most serious causes<br />

of mycosis in man. Both individuals with a<br />

healthy immune system and, more frequently,<br />

immunocompromised patients such as AIDS<br />

sufferers or organ transplant patients, can be<br />

attacked. There is a vast amount of literature<br />

on C. neoformans, including an authoritative<br />

monographic treatment (Casadevall & Perfect,<br />

1998) and several good reviews (e.g. Kwon-Chung<br />

& Bennett, 1992; Mitchell & Perfect, 1995;<br />

Buchanan & Murphy, 1998).<br />

Varieties of C. neoformans and their habitats<br />

One important feature of C. neoformans is the<br />

coat of mucilage which surrounds actively<br />

growing yeast cells. There are four mucilage<br />

serotypes, A D, and these have been correlated<br />

with two varieties of C. neoformans, i.e. var. gattii<br />

(serotypes B and C) and var. neoformans (A and D).<br />

Franzot et al. (1999) have put forward evidence<br />

that the latter should be separated in<strong>to</strong> var.<br />

neoformans (D) and var. grubii (A). Some serotypes<br />

cannot be assigned clearly <strong>to</strong> any one variety<br />

(e.g. serotype AD). Although the three varieties<br />

probably diverged some 18 37 million years<br />

ago, they still show occasional hybridization in<br />

nature and the ability <strong>to</strong> mate in the labora<strong>to</strong>ry.<br />

Many of the hybrids isolated from nature are of<br />

recent origin, and there is evidence that human<br />

activity has brought <strong>to</strong>gether strains which were<br />

previously living in geographic isolation. Thus,<br />

the species C. neoformans may be undergoing<br />

de-diversification at present (Xu et al., 2000).<br />

The anamorph teleomorph connection is such<br />

that C. neoformans var. neoformans and var. grubii<br />

correspond <strong>to</strong> F. neoformans var. neoformans,<br />

whereas the teleomorph of C. neoformans var.<br />

gattii is F. neoformans var. bacillispora which differs


662 BASIDIOMYCETE YEASTS<br />

in that its basidiospores are rod-shaped rather<br />

than spherical or ellipsoid.<br />

As with most if not all fungi pathogenic <strong>to</strong><br />

man, the ability of C. neoformans <strong>to</strong> cause disease<br />

is serendipi<strong>to</strong>us, and the human body represents<br />

a dead-end in the life cycle of the pathogen.<br />

Disease outbreaks are preceded by contact of<br />

humans with the fungus in its natural habitat,<br />

but the precise identity of this has been difficult<br />

<strong>to</strong> track down. Ellis and Pfeiffer (1990a) noted the<br />

geographic co-occurrence of human cases caused<br />

by C. neoformans var. gattii and the distribution of<br />

Eucalyptus camaldulensis trees in Australia, and<br />

were able <strong>to</strong> show that the fungus is associated<br />

with the flowers, bark and litter of Eucalyptus.<br />

In other countries, different plants may also<br />

act as hosts. Further, clinical and environmental<br />

isolates of C. neoformans var. gattii in Australia<br />

have been shown <strong>to</strong> be genetically identical<br />

(Sorrell et al., 1996). <strong>Third</strong>ly, the confinement of<br />

C. neoformans var. gattii <strong>to</strong> rural areas may explain<br />

why its abundance has barely increased in the<br />

wake of the AIDS epidemic, in marked contrast<br />

<strong>to</strong> var. neoformans and especially var. grubii which<br />

accounts for 99% of infections in AIDS patients<br />

(Mitchell & Perfect, 1995). Both these latter<br />

varieties are common in urban habitats, being<br />

associated with the nests and droppings of birds,<br />

especially pigeons (Emmons, 1955) but also caged<br />

birds. Although birds certainly spread the<br />

fungus, they are more likely <strong>to</strong> be vec<strong>to</strong>rs than<br />

primary hosts in nature because they are not<br />

themselves affected by cryp<strong>to</strong>coccosis. As yet<br />

unidentified plants are probably the primary<br />

substratum, and Lazéra et al. (1996) have shown<br />

that C. neoformans var. neoformans (presumably<br />

including var. grubii) occurs in the hollows of<br />

living tree trunks.<br />

Whereas C. neoformans var. gattii, like<br />

Eucalyptus camaldulensis, is confined <strong>to</strong> tropical<br />

and subtropical areas, vars. neoformans and grubii<br />

are cosmopolitan.<br />

Life cycle<br />

The life cycle of F. neoformans is unusual in<br />

several respects (Kwon-Chung, 1998) as shown in<br />

Fig. 24.2. The fungus is heterothallic with a bipolar<br />

mating system (two mating types, a and a).<br />

The yeast cells are haploid and uninucleate,<br />

and they reproduce by budding. Conjugation<br />

between two cells of compatible mating types<br />

gives rise <strong>to</strong> a limited dikaryotic mycelium with<br />

septa bearing clamp connections. From this,<br />

elongated metabasidia arise and terminate in a<br />

swollen tip. Nuclear fusion occurs in the basidial<br />

stalk, and meiosis followed by repeated mi<strong>to</strong>tic<br />

divisions in the swollen apex. This produces four<br />

patches which bud off a chain of basidiospores,<br />

each of which contains a single haploid nucleus.<br />

Meanwhile, mi<strong>to</strong>sis continues in the basidial<br />

cy<strong>to</strong>plasm. Each patch can produce numerous<br />

basidiospores so that a column consisting of four<br />

chains is formed. The migration of nuclei in<strong>to</strong><br />

the basidiospores is random so that each chain<br />

may contain basidiospores of both mating types.<br />

This kind of basidium is unique <strong>to</strong> F. neoformans.<br />

‘Self-fertility’ is a frequently observed<br />

phenomenon in a mating type strains of<br />

F. neoformans and involves the production of a<br />

monokaryotic mycelium with incomplete clamp<br />

connections, giving rise <strong>to</strong> basidia and basidiospores<br />

presumably without meiosis (Wickes et al.,<br />

1996). These should therefore be called conidia.<br />

Of all varieties of C. neoformans, strains containing<br />

mating type a are isolated much more<br />

frequently than a strains from natural situations<br />

as well as from patients (Kwon-Chung & Bennett,<br />

1978), and the ability of a but not a strains <strong>to</strong><br />

form dry wind-dispersed conidia may explain<br />

their greater abundance in nature. Kwon-Chung<br />

et al. (1992), working on strains of C. neoformans<br />

var. neoformans which were genetically identical<br />

<strong>to</strong> each other except for their mating type, found<br />

that the a strain was also more virulent than the<br />

a strain. The conclusion was that the a idiomorph<br />

might encode virulence fac<strong>to</strong>rs.<br />

Both a and a mating type idiomorphs of<br />

C. neoformans var. neoformans and var. grubii have<br />

now been sequenced (Lengeler et al., 2002) and<br />

were found <strong>to</strong> be unusually large (105 130 kb).<br />

Approximately 20 genes are encoded by the<br />

entire sequence, including pheromones, recep<strong>to</strong>rs,<br />

transcription fac<strong>to</strong>rs and components of<br />

signalling cascades. Several genes are unique <strong>to</strong><br />

either a or a, and even those common <strong>to</strong> both are<br />

arranged in different positions, making recombination<br />

within the mating type region impossible.<br />

Lengeler et al. (2002) have likened the


HETEROBASIDIOMYCETE YEASTS<br />

663<br />

Fig 24.2 Life cycle of Filobasidiella neoformans. Note that only the a but not the a mating type is able <strong>to</strong> reproduce by haploid<br />

conidia. Based partly on Kwon-Chung and Bennett (1992) and Kwon-Chung (1998).Open and closed circles represent haploid<br />

nuclei of opposite mating type; diploid nuclei are larger and half-filled. Key events in the life cycle are plasmogamy (P), karyogamy<br />

(K) and meiosis (M).<br />

mating type idiomorph of C. neoformans <strong>to</strong> the sex<br />

chromosomes of other organisms. This situation<br />

is most unusual for Basidiomycota which typically<br />

have two unlinked mating type loci (e.g.<br />

A and B in Coprinus cinereus or Schizophyllum<br />

commune; see p. 508) encoding a recep<strong>to</strong>r/pheromone<br />

system and a transcription fac<strong>to</strong>r.<br />

Another, though structurally different, example<br />

of a sex chromosome may be found in Ustilago<br />

hordei (see p. 637).<br />

Infection and therapy<br />

The available evidence suggests that infection<br />

with C. neoformans is initiated by inhalation, and<br />

that the disease is not spread between humans.<br />

This is probably because the infectious particles<br />

must be very small (


664 BASIDIOMYCETE YEASTS<br />

5-fluorocy<strong>to</strong>sine over several weeks is effective if<br />

the disease has been diagnosed sufficiently early<br />

(Mitchell & Perfect, 1995). For AIDS sufferers, ‘the<br />

therapeutic goal is <strong>to</strong> ablate the symp<strong>to</strong>ms and<br />

signs of cryp<strong>to</strong>coccosis until the patient dies of<br />

other causes’ (Kwon-Chung & Bennett, 1992).<br />

Antifungal drugs, especially fluconazole, may<br />

have <strong>to</strong> be administered for the rest of the<br />

patient’s life because of the high risk of relapses.<br />

Cryp<strong>to</strong>coccosis, like certain other fungal<br />

diseases, is an AIDS-defining illness, i.e. patients<br />

diagnosed as HIV-positive and suffering from<br />

cryp<strong>to</strong>coccosis are considered <strong>to</strong> have developed<br />

AIDS.<br />

Virulence fac<strong>to</strong>rs<br />

Much work is being done <strong>to</strong> identify the properties<br />

which enable C. neoformans <strong>to</strong> cause disease<br />

(reviewed by Buchanan & Murphy, 1998;<br />

Casadevall & Perfect, 1998; Perfect, 2004). The<br />

most important and obvious virulence fac<strong>to</strong>r is<br />

the capsule which surrounds actively growing<br />

yeast cells of C. neoformans and protects them<br />

from adverse environmental effects. It also seems<br />

<strong>to</strong> inhibit the phagocy<strong>to</strong>sis of yeast cells by<br />

macrophages. The capsule can be visualized by<br />

light microscopy of yeast cells mounted in India<br />

ink and can be rather more substantial than that<br />

shown for Rhodo<strong>to</strong>rula in Fig. 24.1b. The major<br />

capsule polysaccharide of C. neoformans is a linear<br />

a-(1,3)-mannan chain, of which roughly every<br />

third mannose moiety is substituted with a<br />

single b-(1,2)-glucuronic acid unit. The presence<br />

of xylose determines the antigenic properties of<br />

the capsule; one (serotype D), two (A), three (B) or<br />

four (C) xylose residues may be present for every<br />

three mannose moieties (Cherniak & Sundstrom,<br />

1994). The capsule polysaccharides are produced<br />

in such profusion that they can be detected in<br />

the blood serum and other body fluids of<br />

infected patients, and this is an important<br />

diagnostic <strong>to</strong>ol (Kwon-Chung & Bennett, 1992).<br />

Mutants unable <strong>to</strong> produce a capsule in the host<br />

are apathogenic.<br />

Both C. neoformans var. gattii and vars. grubii/<br />

neoformans have been shown <strong>to</strong> undergo switches<br />

in colony phenotype from smooth <strong>to</strong> mucoid,<br />

wrinked or pseudohyphal. Switching between<br />

smooth and mucoid appears <strong>to</strong> be readily<br />

reversible in the mammalian host in var. gattii,<br />

but less so in the latter two. The mucoid type is<br />

characterized by an increase in the thickness of<br />

the polysaccharide capsule. In C. neoformans var.<br />

neoformans and var. grubii a thick capsule coincides<br />

with an enhanced resistance <strong>to</strong> antimycotics,<br />

and mucoid strains may be selected by<br />

prolonged chemotherapy (Guerrero et al., 2006).<br />

In mice infected with C. neoformans var. gattii, Jain<br />

et al. (2006) have shown that mucoid forms are<br />

associated preferentially with pulmonary infections,<br />

presumably because of an enhanced<br />

resistance <strong>to</strong> intracellular digestion by phagocytes.<br />

Smooth cells with their thinner coat were<br />

preferentially isolated from infected brain tissue.<br />

In contrast <strong>to</strong> morphotype switching in Candida<br />

albicans (p. 277), no sexual function has been<br />

suggested for switching in C. neoformans.<br />

A second important virulence fac<strong>to</strong>r is melanin.<br />

The ability of C. neoformans <strong>to</strong> synthesize<br />

melanin distinguishes this species from other<br />

members of the genus, and the melanized cell<br />

wall has been suggested <strong>to</strong> protect the cell against<br />

oxidative stress such as that encountered during<br />

the oxidative burst after ingestion by macrophages<br />

(Wang et al., 1995). The pathway of<br />

melanin biosynthesis in C. neoformans is different<br />

from the dihydroxynaphthalene (DHN) route<br />

found in most fungi (see Fig. 12.46). The precursor<br />

molecule is 3,4-dihydroxyphenylalanine (DOPA)<br />

which cannot be synthesized by C. neoformans<br />

but, if present, can be oxidized <strong>to</strong> quinones by a<br />

laccase-type enzyme, and these quinones spontaneously<br />

polymerize <strong>to</strong> melanin (Salas et al., 1996).<br />

Catecholamines such as DOPA are present at high<br />

levels in the brain, and this may explain the<br />

preferential accumulation of C. neoformans in<br />

brain tissue (Polachek et al., 1990).<br />

A third important fac<strong>to</strong>r is, of course, the<br />

ability of C. neoformans <strong>to</strong> grow at 37 39°C,<br />

which is unique among Cryp<strong>to</strong>coccus spp. Several<br />

gene products are required for growth at 37°C,<br />

and prominent among them is calcineurin,<br />

a Ca 2þ -regulated serine/threonine phosphatase<br />

involved in eukaryotic cellular signalling<br />

(Odom et al., 1997). Calcineurin is the target of<br />

a complex formed between a cyclophilin protein<br />

involved in protein folding, and cyclosporin A<br />

(see Fig. 12.24a), an immunosuppressive drug


HETEROBASIDIOMYCETE YEASTS<br />

665<br />

widely used <strong>to</strong> prevent the graft rejection<br />

reaction after organ transplantations. In vitro,<br />

growth of C. neoformans is inhibited by cyclosporin<br />

A at 37°C but not at lower temperatures.<br />

The use of cyclosporin A for human antifungal<br />

therapy is impossible because its immunosuppressive<br />

activity outweighs the antifungal effect.<br />

However, Cruz et al. (2000) have reported cyclosporin<br />

derivatives which are antifungal but do<br />

not have immunosuppressive properties, i.e. they<br />

appear <strong>to</strong> interfere with fungal calcineurin<br />

signalling but not with the equivalent human<br />

signalling chain involved in the immune<br />

response. It remains <strong>to</strong> be seen whether these<br />

substances are sufficiently specific for use as new<br />

antifungal drugs.<br />

24.2.2 Phaffia and Xanthophyllomyces<br />

Phaffia rhodozyma and its putative teleomorph<br />

Xanthophyllomyces dendrorhous have aroused<br />

considerable interest because they are among<br />

very few fungi producing the commercially<br />

valuable carotenoid pigment astaxanthin (see<br />

Fig. 24.8), and probably the only ones <strong>to</strong> do so in<br />

pure culture. Astaxanthin is of importance in the<br />

fish farming industry because salmonid fish<br />

(salmon and trout) require a minimum level of<br />

astaxanthin in their food for healthy growth.<br />

One reason for this may be that these fish<br />

contain high levels of polyunsaturated fatty<br />

acids which are susceptible <strong>to</strong> peroxidation by<br />

reactive oxygen species such as the hydroxyl<br />

radical (HO) or superoxide radical (O 2 ).<br />

Astaxanthin and the related pigment canthaxanthin<br />

are also responsible for the orange<br />

pigmentation of salmon steak. In nature, astaxanthin<br />

travels the food chain phy<strong>to</strong>plank<strong>to</strong>n !<br />

zooplank<strong>to</strong>n ! larger crustaceans ! salmon.<br />

Simple methods <strong>to</strong> extract and analyse astaxanthin<br />

from Phaffia and salmon have been<br />

described by Weber and Davoli (2003). In view<br />

of the correlation between oxidative processes<br />

and cancer or degenerative diseases related <strong>to</strong><br />

ageing, astaxanthin is also gaining popularity<br />

as a ‘nutraceutical’, i.e. an additive <strong>to</strong> the<br />

human diet.<br />

Astaxanthin is now produced commercially<br />

by <strong>to</strong>tal chemical synthesis, by extracting it from<br />

the exoskele<strong>to</strong>ns of shellfish, or by microbial<br />

fermentation using Phaffia or the alga<br />

Haema<strong>to</strong>coccus pluvialis. Whereas wild-type strains<br />

of Phaffia synthesize a limited amount of astaxanthin<br />

(typically less than 300 mgg 1 dry<br />

weight), the astaxanthin levels in strains used<br />

for industrial production are at least 10-fold<br />

higher. In-depth reviews of biotechnological<br />

aspects of astaxanthin production have been<br />

written by Johnson and An (1991) and Johnson<br />

and Schroeder (1995a,b).<br />

It is unclear why, among the numerous red<br />

yeasts, Phaffia and Xanthophyllomyces are the only<br />

ones as yet known <strong>to</strong> synthesize astaxanthin.<br />

Part of the answer may lie in their unusual<br />

habitat, all strains known <strong>to</strong> date having been<br />

isolated from the slime fluxes of broad-leaved<br />

trees, especially birch (Betula spp.) in Alaska,<br />

Japan, Scandinavia and Russia (Phaff, 1990).<br />

Schroeder and Johnson (1995) have presented<br />

evidence that birch sap contains a pho<strong>to</strong>sensitizer,<br />

i.e. a substance that becomes energized by<br />

UV light. When this passes on its excitation<br />

energy, it can transform ground-state oxygen<br />

(triplet oxygen,<br />

3 O 2 ) <strong>to</strong> the reactive singlet<br />

oxygen ( 1 O 2 ) state. This can be returned <strong>to</strong> its<br />

ground state under dissipation of the excess<br />

energy as heat by astaxanthin and other carotenoids.<br />

Astaxanthin appears <strong>to</strong> be necessary for<br />

the survival of Phaffia under the oxidizing<br />

conditions prevalent in birch sap. In fact,<br />

cultivation and mutant studies have shown<br />

that astaxanthin is able <strong>to</strong> protect its producing<br />

organism against a wide range of oxidative<br />

stresses caused by H 2 O 2 , singlet oxygen and the<br />

hydroxyl and superoxide radicals, and that<br />

astaxanthin biosynthesis is increased under<br />

such conditions (Schroeder & Johnson, 1993).<br />

In contrast, b-carotene is the main pigment at<br />

low oxygen partial pressure, i.e. Phaffia cultures<br />

grown under reduced aeration appear yellow<br />

instead of orange (see Fig. 24.8).<br />

The original isolate of Phaffia is a purely<br />

asexual strain, but other isolates show sexual<br />

reproduction under suitable conditions characterized<br />

by nitrogen starvation and the presence<br />

of polyols (Kucsera et al., 1998). This perfect state<br />

was named Xanthophyllomyces dendrorhous, and<br />

there is still disagreement as <strong>to</strong> whether Phaffia


666 BASIDIOMYCETE YEASTS<br />

and Xanthophyllomyces represent the same or<br />

closely related species (Fell & Blatt, 1999;<br />

Kucsera et al., 2000). Xanthophyllomyces appears<br />

<strong>to</strong> be homothallic, vegetative cells being diploid.<br />

Mating can occur between a mother cell and its<br />

bud, giving rise <strong>to</strong> a tetraploid zygote (Kucsera<br />

et al., 1998). A long thin aseptate metabasidium is<br />

formed, and the fusion nucleus migrates <strong>to</strong><br />

the tip of it, undergoing meiosis in the course<br />

of the journey. About 2 7 diploid basidiospores<br />

are produced and released passively (Fig. 24.4c).<br />

Vegetative reproduction is by budding (Fig.<br />

24.4a), but the nuclear events are unusual in<br />

that the mother nucleus migrates in<strong>to</strong> the bud<br />

and divides there, followed by the return of one<br />

of the daughter nuclei <strong>to</strong> the mother cell<br />

(Slaninova et al., 1999). Thick-walled chlamydospores<br />

(Fig. 24.4b) are occasionally formed, and<br />

these germinate by mi<strong>to</strong>tic budding, i.e. they are<br />

not equivalent <strong>to</strong> the teliospores of other<br />

basidiomycete yeasts. The life cycle of<br />

Xanthophyllomyces rhodozyma is summarized in<br />

Fig. 24.3. In general, there is a considerable<br />

variation between different isolates; for instance,<br />

the number of chromosomes ranges from 7 <strong>to</strong> 17<br />

(Kucsera et al., 2000).<br />

24.3 Urediniomycete yeasts<br />

Yeasts belonging <strong>to</strong> the Urediniomycetes appear<br />

<strong>to</strong> fall in<strong>to</strong> four phylogenetic groups (see<br />

Table 24.1), some of which still need <strong>to</strong> be<br />

integrated in<strong>to</strong> higher taxa (families and<br />

orders). Many species contain carotenoids, i.e.<br />

they are red yeasts. Some of them, especially<br />

species belonging <strong>to</strong> the Sporodiobolus clade<br />

(provisional order Sporidiales), are extremely<br />

common in the environment. For this reason,<br />

we shall consider them here. There are two<br />

important genera in the Sporidiales, Sporidiobolus<br />

(anamorph Sporobolomyces) and Rhodosporidium<br />

(anamorph Rhodo<strong>to</strong>rula).<br />

24.3.1 Sporidiales<br />

In addition <strong>to</strong> budding, Sporobolomyces spp. also<br />

form asexual spores which are ejected from a<br />

sterigma in<strong>to</strong> the air by the surface-tension<br />

catapult mechanism and are therefore called<br />

ballis<strong>to</strong>conidia. If an agar culture is incubated<br />

upside down, a mirror image of the colony will<br />

be deposited as spores on the Petri dish lid. For<br />

this reason, Sporobolomyces is known as a mirror<br />

yeast. One of the most frequently encountered<br />

species, S. roseus, is now known <strong>to</strong> represent a<br />

Fig 24.3 Life cycle of<br />

Xanthophyllomyces dendrorhous.<br />

Vegetative cells are diploid and<br />

reproduce by budding.The fungus<br />

is homothallic, and fusion between<br />

mother and daughter cell establishes<br />

a tetraploid zygote which forms an<br />

elongate aseptate basidium in which<br />

meiosis occurs.The diploid<br />

basidiospores germinate by budding.<br />

Xanthophyllomyces can survive<br />

unfavourable conditions by means<br />

of thick-walled chlamydospores.<br />

Half-filled circles represent diploid<br />

nuclei; tetraploid nuclei are larger and<br />

divided in<strong>to</strong> four segments. Key events<br />

in the life cycle are plasmogamy (P),<br />

karyogamy (K) and meiosis (M).


UREDINIOMYCETE YEASTS<br />

667<br />

Fig 24.4 Microscopic features of Phaffia rhodozyma (a) and<br />

Xanthophyllomyces (b,c), a closely related teleomorph.<br />

(a) Vegetatively dividing yeast cells. (b) Thick-walled<br />

chlamydospore in the process of germination bybudding about<br />

3 h after transfer <strong>to</strong> a fresh medium. (c) Aseptate elongate<br />

basidia of Xanthophyllomyces sp. producing basidiospores at<br />

their tips. All images <strong>to</strong> same scale.<br />

complex of several species which are difficult <strong>to</strong><br />

distinguish by any means other than DNA<br />

sequencing (Bai et al., 2002). Rhodo<strong>to</strong>rula spp.<br />

also reproduce vegetatively by budding but are<br />

unable <strong>to</strong> produce ballis<strong>to</strong>conidia. Apart from<br />

this difference, the two genera are very close <strong>to</strong><br />

each other, overlapping in phylogenetic trees<br />

and also possessing similar life cycles.<br />

Life cycle<br />

The budding phase of Sporobolomyces roseus is<br />

uninucleate, and nuclear division occurs at the<br />

time of bud formation (Buller, 1933). If ballis<strong>to</strong>conidia<br />

are <strong>to</strong> be formed, a conical sterigma<br />

develops vertically and bears an asymmetric<br />

spore which strongly resembles a basidiospore<br />

(Figs. 24.1e and 24.5). A daughter nucleus passes<br />

in<strong>to</strong> this spore. Following the fusion of Buller’s<br />

drop with the adaxial blob, the spore is flicked<br />

away for a distance of about 0.1 mm. Colonies of<br />

S. roseus begin <strong>to</strong> form ballis<strong>to</strong>conidia after<br />

2 3 days on most agar media, and they are<br />

recognized by the presence of numerous satellite<br />

colonies outside of the margin of the parent<br />

colony. A single sterigma may produce a second<br />

or even a third spore, and occasionally two or<br />

three sterigmata arise from one cell.<br />

Although S. roseus can form pseudohyphae<br />

and true hyphae, no sexual reproduction has<br />

been observed. However, a related species,<br />

Sporidiobolus (Sporobolomyces, Aessosporon) salmonicolor<br />

(formerly also called S. odorus), does undergo<br />

the full sexual cycle (Fig. 24.7). This species is<br />

heterothallic with a bipolar (unifac<strong>to</strong>rial) mating<br />

system (Fell & Statzell-Tallman, 1981). Following<br />

conjugation between compatible haploid uninucleate<br />

cells, a dikaryotic mycelium with clamp<br />

connections develops and eventually produces<br />

globose binucleate teliospores (Fig. 24.6; Bandoni<br />

et al., 1971). Karyogamy follows. Meiosis occurs<br />

during teliospore germination, which gives<br />

rise <strong>to</strong> an aseptate metabasidium. This buds<br />

off haploid monokaryotic basidiospores at its<br />

tip. Life cycles of this kind have also been<br />

described in some other members of the<br />

Sporidiales, with minor variations such as<br />

the presence of transverse septa in the metabasidia,<br />

or the absence of ballis<strong>to</strong>conidia (see<br />

Fig. 24.1f).<br />

There is, however, a complication in the life<br />

cycle of S. salmonicolor because, according <strong>to</strong><br />

van der Walt (1970), yeast cells may be haploid<br />

or diploid. The diploid cells are larger than the<br />

haploid ones. Both types are capable of reproducing<br />

by budding and by producing ballis<strong>to</strong>conidia,<br />

but only the diploid cells can additionally<br />

develop directly in<strong>to</strong> thick-walled teliospores.<br />

These germinate by means of a short aseptate<br />

promycelium producing yeast-like basidiospores<br />

at its tip (Fig. 24.7).<br />

According <strong>to</strong> Fell and Statzell-Tallman (1998),<br />

a similar life cycle is found in Rhodosporidium<br />

<strong>to</strong>ruloides (anamorph Rhodo<strong>to</strong>rula glutinis). In this<br />

species, diploid yeast cells are thought <strong>to</strong> arise if<br />

there is a failure of meiosis during teliospore<br />

germination. In contrast <strong>to</strong> the observations by


668 BASIDIOMYCETE YEASTS<br />

Fig 24.5 Sporobolomycesroseus.(a c)<br />

Various stages in the budding of cells.<br />

(d) Cell bearing a sterigma and a<br />

ballis<strong>to</strong>conidium. (e) Cell with two<br />

sterigmata. (f) Cell with three sterigmata.<br />

van der Walt (1970) on Sporidiobolus, the diploid<br />

yeast cells of R. <strong>to</strong>ruloides do not develop in<strong>to</strong><br />

teliospores directly, but first produce a mycelium<br />

which in turn forms teliospores. Conjugation<br />

between haploid cells of compatible mating<br />

type is mediated by peptide hormones. One of<br />

them, rhodo<strong>to</strong>rucine A, has been identified as<br />

an undecapeptide with a farnesyl side chain<br />

(Kamiya et al., 1979).<br />

Some species of Rhodosporidium and Sporidiobolus<br />

are entirely homothallic, and here all yeast<br />

cells seem <strong>to</strong> be diploid (Fell & Statzell-Tallman,<br />

1998; Statzell-Tallman & Fell, 1998). Meiosis<br />

in a diploid yeast cell establishes a clamped<br />

dikaryotic mycelium which produces teliospores.<br />

Germinating teliospores give rise <strong>to</strong> metabasidia<br />

which produce diploid basidiospores.<br />

Ecology<br />

Sporobolomyces, Rhodo<strong>to</strong>rula and other yeasts<br />

belonging <strong>to</strong> the Urediniomycetes are found in<br />

diverse habitats such as the soil (Sláviková &<br />

Vadkertiová, 2003) and the sea, including deep<br />

sea locations (Nagahama et al., 2001). However,<br />

their most prominent habitat is healthy and<br />

moribund vegetation (Nakase, 2000) from which<br />

they can be isolated throughout the year.<br />

Together with the black yeast Aureobasidium<br />

(Ascomycota; see p. 484), they form a major<br />

component of the phylloplane yeast population.<br />

Breeze and Dix (1981) have estimated that<br />

yeasts (including ascomycetes) produce up <strong>to</strong><br />

50 times more biomass than hyphal fungi<br />

on Acer leaves throughout the growing season.<br />

Sporobolomyces roseus is the most abundant phylloplane<br />

yeast, comprising for example 76% of the<br />

yeast population on grapes (de la Torre et al.,<br />

1999). Scanning electron microscopy studies have<br />

shown that cells of S. roseus may form sheets<br />

of mucilage by which they adhere <strong>to</strong> the leaf<br />

surface. There is no evidence that the growth of<br />

the yeasts causes corrosion of the cuticle (Bashi &<br />

Fokkema, 1976). There is interest in the suggestion<br />

that Sporobolomyces or Rhodo<strong>to</strong>rula on leaf<br />

surfaces may compete with foliar pathogens,<br />

and that biological control of the pathogens<br />

might be possible (Fokkema & van der Meulen,<br />

1976). Such an approach is especially promising<br />

in the control of post-harvest diseases<br />

because the incubation conditions can be more<br />

precisely controlled (Janisiewicz & Bors, 1995;<br />

Janisiewicz & Korsten, 2002). Many Sporobolomyces


UREDINIOMYCETE YEASTS<br />

669<br />

Fig 24.6 Sporidiobolus salmonicolor.<br />

(a) Budding cell. (b) Cell with<br />

sterigmata. (c) Conjugation of<br />

haploid cells and initiation of a<br />

dikaryotic hypha. (d) Young<br />

dikaryotic hyphae, the original<br />

conjugants obscured.The arrow<br />

indicates the first clamp<br />

connection. (e) Late stage of<br />

chlamydospore development;<br />

the paired nuclei are visible.<br />

After Bandoni et al.(1971).<br />

and Rhodo<strong>to</strong>rula spp. appear <strong>to</strong> be associated with<br />

lesions caused by plant pathogens such as rust<br />

fungi, or by parasites such as nema<strong>to</strong>des.<br />

Ballis<strong>to</strong>conidia of Sporobolomyces are frequent<br />

in the air, especially during warm summer<br />

nights, and concentrations of these spores<br />

may reach values of up <strong>to</strong> 10 6 m 3 (Gregory &<br />

Sreeramulu, 1958). Such high concentrations are<br />

a cause for concern because Sporobolomyces has<br />

been shown <strong>to</strong> be a respira<strong>to</strong>ry allergen (Evans,<br />

1965).<br />

Phylloplane yeast populations occupy an<br />

exposed habitat and are therefore susceptible<br />

<strong>to</strong> environmental changes. Sporobolomyces roseus<br />

has been proposed as an indica<strong>to</strong>r of air quality,<br />

low colony counts correlating with heavy air<br />

pollution (Dowding & Richardson, 1990). The<br />

results can be superimposed on those obtainable


670 BASIDIOMYCETE YEASTS<br />

Fig 24.7 Possible life cycle of Sporidiobolussalmonicolor.Both diploid stages (large split nuclei) and haploid stages (small black or white<br />

nuclei) exist and can reproduce by budding and ballis<strong>to</strong>conidium formation. Plasmogamy of two yeast cells of compatible mating<br />

types or conversion of a diploid yeast cell leads <strong>to</strong> the formation of a teliospore which undergoes meiosis upon germination.<br />

Failure of meiosis in the metabasidium or conjugation between two haploid yeast cells could give rise <strong>to</strong> diploid yeast cells.<br />

Key events in the life cycle are plasmogamy (P), karyogamy (K) and meiosis (M).<br />

with the biomoni<strong>to</strong>ring of lichens (see p. 454),<br />

and both groups of organisms are susceptible<br />

<strong>to</strong> SO 2 .<br />

Phylloplane yeasts including S. roseus are also<br />

sensitive <strong>to</strong> UV radiation, and enhanced irradiation<br />

leads <strong>to</strong> a decline in their abundance on<br />

leaves (Newsham et al., 1997). However, some<br />

protection may be afforded by the carotenoids<br />

produced by red yeasts, since carotenoid-deficient<br />

mutants are more sensitive <strong>to</strong> UV light<br />

than the red wild-type strains. Pure-culture<br />

studies with Sporobolomyces and Rhodo<strong>to</strong>rula spp.<br />

often show enhanced production of carotenoids<br />

under oxidative stress such as high aeration or<br />

the addition of radical genera<strong>to</strong>rs. Further, the<br />

carotenoid spectrum may also change; if it does,<br />

the shift is commonly from g- and/or b-carotene<br />

under reduced oxygen pressure <strong>to</strong> <strong>to</strong>rulene<br />

and oxidized carotenoids (e.g. <strong>to</strong>rularhodin)<br />

at oxidative stress (Fig. 24.8; Sakaki et al., 2002;<br />

Davoli et al., 2004). This may represent an<br />

adaptive strategy because many of the oxygencontaining<br />

carotenoids (i.e. xanthophylls) have<br />

superior anti-oxidant properties as compared<br />

<strong>to</strong> b-carotene (Martin et al., 1999). Other metabolites<br />

(e.g. vitamin E) or enzymes (superoxide<br />

dismutase, catalase) with anti-oxidant activities<br />

may also be produced by red yeasts so that<br />

the correlation between carotenoid production<br />

and anti-oxidant protection is not always<br />

absolute.<br />

24.4 Ustilaginomycete yeasts<br />

Most yeasts belonging <strong>to</strong> the Ustilaginomycetes<br />

are the monokaryotic stages of smut fungi<br />

(Section. 23.2) or of Exobasidium (Section 23.4)<br />

and live saprotrophically on plant surfaces.


USTILAGINOMYCETE YEASTS<br />

671<br />

Fig 24.8 Examples of the effect of aeration on carotenoid production in red yeasts belonging <strong>to</strong> the Basidiomycota. In most red<br />

yeasts, carotenoid biosynthesis progresses via lycopene at least <strong>to</strong> the first end-group cyclization <strong>to</strong> give g-carotene. Several yeasts<br />

adjust their carotenoid spectrum in response <strong>to</strong> the level of oxidative stress in their environment. In Sporobolomyces and Rhodo<strong>to</strong>rula<br />

spp. (Urediniomycetes), g-carotene and/or the bicyclic b-carotene accumulate under microaerophilic conditions (dotted arrow),<br />

whereas <strong>to</strong>rulene and oxygen-containing carotenoids (xanthophylls) such as <strong>to</strong>rularhodin are produced under oxidative stress<br />

(solid arrows). In contrast, in Phaffia rhodozyma (Heterobasidiomycetes) the b-carotene produced at low partial oxygen pressure<br />

is directly oxidized <strong>to</strong> bicyclic xanthophylls (notably astaxanthin) under oxidative stress.<br />

Here we shall briefly consider a third group<br />

of Ustilaginomycetes, the Malasseziales, which<br />

colonize the skin of warm-blooded animals,<br />

i.e. mammals and birds. Good accounts of<br />

this order have been written by Ashbee and<br />

Evans (2002), Crespo Erchiga and Delgado<br />

Florencio (2002), Inamadar and Palit (2003) and<br />

Batra et al. (2005).<br />

24.4.1 Malasseziales<br />

This group has been named after the French<br />

pathologist Louis C. Malassez who was one of the<br />

pioneers in this field. Following an eventful<br />

taxonomic his<strong>to</strong>ry, there is now only one<br />

genus, Malassezia (formerly Pityrosporum) with<br />

seven species (Guillot & Guého, 1995; Guého<br />

et al., 1996). All of them are haploid without<br />

known sexual stages.<br />

The yeast cells show several unusual characteristics<br />

by which they can be recognized.<br />

Budding is enteroblastic, with the daughter cell<br />

arising in a unipolar fashion from an exceptionally<br />

broad ring-like bud scar of the mother cell<br />

(Guého & Meyer, 1989). The ultrastructure of<br />

Malassezia is most unusual. The cell wall is thick,<br />

with its inner surface sculptured in<strong>to</strong> spiral<br />

ridges (Fig. 24.9) which cast the plasma<br />

membrane in<strong>to</strong> corresponding grooves (David<br />

et al., 2003). There is considerable morphological<br />

plasticity in some Malassezia spp., notably<br />

M. furfur, in that the shape of yeast cells may<br />

change from globose <strong>to</strong> elongated after repeated<br />

subculturing. Hyphae may also be formed under<br />

certain conditions.<br />

All but one species (M. pachydermatis) are<br />

obligately lipophilic and are isolated from skin<br />

samples if standard agar media are overlaid with<br />

a thin film of olive oil. It is possible <strong>to</strong> isolate<br />

most Malassezia spp. from the skin of humans<br />

and other animals, with M. pachydermatis


672 BASIDIOMYCETE YEASTS<br />

Fig 24.9 Malassezia pachydermatis.<br />

Freeze fracture image of a budding cell<br />

showing the inner surface of the wall.<br />

The spiral ridges are not seen in the collar<br />

region. From David et al. (2003), by copyright<br />

permission of Scripta Media,Brno.Original<br />

image kindly provided by M. David, M.Gabriel<br />

and M. Kopecka¤.<br />

showing a predilection for dogs and cats where it<br />

can cause skin infections (Guillot & Bond, 1999).<br />

Most healthy humans carry Malassezia spp., but<br />

the species composition and density of colonization<br />

are altered in patients affected by dermatitis<br />

(Sugita et al., 2001; Crespo Erchiga & Delgado<br />

Florencio, 2002). The pattern of colonization is<br />

dependent upon several fac<strong>to</strong>rs, e.g. the degree of<br />

sweating and the amount of lipid produced.<br />

Malassezia spp. are often most abundant on<br />

people in early adulthood because at that age<br />

the lipid-producing sebaceous glands at the base<br />

of the hair shaft are most active. A clear role<br />

in pathogenesis has been demonstrated only<br />

for M. globosa which causes a superficial mycosis<br />

known as pityriasis versicolor in adults below<br />

middle-age (Gupta et al., 2002). The disease is<br />

more common in the tropics than in cooler<br />

climates. Infected skin areas differ in pigmentation<br />

from the normal skin. Infections are<br />

associated mainly with the hyphal growth form<br />

of M. globosa whereas commensal growth is yeastlike<br />

(Crespo Erchiga & Delgado Florencio, 2002).<br />

Occasionally, Malassezia spp. also cause contaminations<br />

of catheters.<br />

The association of Malassezia spp. with<br />

dandruff has been suggested but is not yet<br />

proven (Piérard-Franchimond et al., 2000). Evidence<br />

in favour of the argument is that dandruff,<br />

like pityriasis versicolor and other superficial<br />

skin infections caused by Malassezia spp., disappears<br />

upon treatment with shampoos or lotions<br />

containing selenium sulphide or other <strong>to</strong>pical<br />

antifungal agents (Kwon-Chung & Bennett,<br />

1992). Further, some Malassezia infections are<br />

more common in immunocompromised patients<br />

than in healthy subjects, and there is evidence<br />

of both humoral and cell-mediated immune<br />

responses against Malassezia in immunocompetent<br />

humans (Ashbee & Evans, 2002).


25<br />

Anamorphic fungi (nema<strong>to</strong>phagous and<br />

aquatic forms)<br />

Throughout this book we have attempted <strong>to</strong><br />

consider fungi showing predominantly or purely<br />

asexual reproduction <strong>to</strong>gether with their known<br />

or suspected teleomorphs. However, certain<br />

groups of taxonomically diverse fungi colonizing<br />

the same specialized habitats or substrates<br />

are best unders<strong>to</strong>od in their ecological context,<br />

especially if they show strikingly similar adaptations<br />

and morphology despite their different<br />

evolutionary his<strong>to</strong>ries. Two cases illustrating<br />

such convergent evolution among anamorphic<br />

fungi are the nema<strong>to</strong>phagous habit and the<br />

aquatic habitat, which we shall consider in turn<br />

in this chapter.<br />

25.1 Nema<strong>to</strong>phagous fungi<br />

Nema<strong>to</strong>des are a very varied group of invertebrates.<br />

They are particularly common as freeliving<br />

saprotrophic species in the soil, around<br />

plant roots, on dung and in all kinds of decomposing<br />

plant matter, as well as in freshwater and<br />

marine habitats. Most saprotrophic nema<strong>to</strong>des<br />

feed on bacteria, although fungal hyphae may<br />

also be consumed. Other species parasitize<br />

animals, releasing their eggs or motile stages<br />

in<strong>to</strong> the environment when their hosts defaecate.<br />

Plant-parasitic species chiefly attack roots as freeliving<br />

or sedentary organisms. Sedentary species<br />

form adult stages inside plant root tissues where<br />

they cause the economically important root knot<br />

diseases (Meloidogyne spp.) or root cyst diseases<br />

(Heterodera spp. and Globodera spp.). Plant-parasitic<br />

nema<strong>to</strong>des are readily recognized because their<br />

mouth parts are modified as stylets with which<br />

they penetrate plant tissues. Gravid females of<br />

cyst nema<strong>to</strong>des enlarge, and their bodies become<br />

converted in<strong>to</strong> a hardened cyst containing the<br />

eggs.<br />

<strong>Fungi</strong> have evolved a range of mechanisms<br />

<strong>to</strong> attack nema<strong>to</strong>des, which can be grouped<br />

in<strong>to</strong> three broad categories described below.<br />

A summary of genera and their taxonomic<br />

relationships is presented in Table 25.1.<br />

Nema<strong>to</strong>phagous fungi are common in most<br />

types of soil which are rich in organic matter,<br />

and they have been found in arctic, temperate<br />

and tropical climates (see Dix & Webster, 1995).<br />

There is no strong evidence of host selectivity.<br />

Preda<strong>to</strong>ry nema<strong>to</strong>phagous fungi produce<br />

a sizeable mycelium in the soil, with trapping<br />

devices formed at intervals along the length<br />

of the hyphae. The varied and occasionally<br />

spectacular trapping mechanisms have captured<br />

the imagination of generations of mycology<br />

students. The most common traps are sticky<br />

knobs, adhesive networks, non-constricting<br />

rings or constricting rings (see pp. 675 680).<br />

Penetration of captured nema<strong>to</strong>des is followed<br />

by the growth of trophic hyphae throughout the<br />

nema<strong>to</strong>de body, and digestion of its contents.<br />

Because most preda<strong>to</strong>ry nema<strong>to</strong>phagous fungi


674 ANAMORPHIC FUNGI<br />

Table 25.1. Examples of the diversity of nema<strong>to</strong>phagous fungi.<br />

Genus Taxonomic affinity Mode of parasitism<br />

Preda<strong>to</strong>ry fungi<br />

Acaulopage, Stylopage Zygomycota Adhesive hyphae<br />

Gamsylella, Dactylellina,<br />

Arthrobotrys (incl. Dactylella,<br />

Dactylaria, Monacrosporium,<br />

Dudding<strong>to</strong>nia)<br />

Drechslerella (incl. Arthrobotrys,<br />

Dactylella, Monacrosporium)<br />

Nema<strong>to</strong>c<strong>to</strong>nus (Hohenbuehelia),<br />

Pleurotus<br />

Endoparasitic fungi<br />

Hap<strong>to</strong>glossa<br />

Orbiliaceae<br />

(Ascomycota)<br />

Orbiliaceae<br />

(Ascomycota)<br />

Pleurotaceae (euagarics<br />

clade,Basidiomycota)<br />

Plasmodiophoromycota<br />

or Oomycota<br />

Adhesive knobs, columns,<br />

nets, non-constricting rings<br />

Constricting rings<br />

Adhesive or poisonous knobs<br />

Gun cells (see Fig. 3.9)<br />

Myzocytium, Nema<strong>to</strong>phthora Oomycota Encysting zoospores<br />

Catenaria Chytridiomycota Encysting zoospores<br />

Harposporium, Drechmeria,<br />

Verticillium, Hirsutella<br />

Clavicipitaceae<br />

(Pyrenomycetes,<br />

Ascomycota)<br />

Ingestion or attachment<br />

of conidia<br />

Egg and cyst parasites<br />

Rhopalomyces Zygomycota Hyphal colonization of eggs<br />

Pochonia chlamydosporia<br />

Paecilomyces lilacinus<br />

Clavicipitaceae<br />

(Pyrenomycetes,<br />

Ascomycota)<br />

Ascomycota (incertae<br />

sedis)<br />

Hyphal colonization of cysts<br />

Hyphal colonization of eggs<br />

have a high saprotrophic potential, including<br />

the ability <strong>to</strong> degrade cellulose, nema<strong>to</strong>des are<br />

probably utilized mainly as a nitrogen supplement<br />

(Barron, 1992).<br />

In the endoparasitic nema<strong>to</strong>phagous fungi<br />

(see p. 680) there is no extensive mycelial<br />

development outside the nema<strong>to</strong>de host, and<br />

these species must therefore be regarded as<br />

obligate parasites in ecological terms. They<br />

exist in the soil as spores which may either<br />

become attached <strong>to</strong> the body of the host,<br />

or become ingested. The spores then germinate<br />

and penetrate the animal, developing a<br />

mycelium within the body of the nema<strong>to</strong>de.<br />

Only reproductive hyphae (conidiophores) penetrate<br />

<strong>to</strong> the outside of the dead colonized<br />

nema<strong>to</strong>de.<br />

Parasites of eggs and cysts (p. 684) are found<br />

in several different taxonomic groups. Typically<br />

these fungi can be isolated from the soil or<br />

rhizosphere as well as from nema<strong>to</strong>de eggs or<br />

cysts, and are therefore viewed as opportunistic<br />

saprotrophs. The most thoroughly studied<br />

representatives are Paecilomyces lilacinus and<br />

Pochonia chlamydosporia (formerly Verticillium<br />

chlamydosporium).


NEMATOPHAGOUS FUNGI<br />

675<br />

We owe much of our knowledge of nema<strong>to</strong>phagous<br />

fungi <strong>to</strong> the numerous publications<br />

by Charles Drechsler. These are cited in many of<br />

the general accounts given by Dudding<strong>to</strong>n<br />

(1955, 1957), Barron (1977, 1981), Dowe (1987),<br />

Gray (1987), Nordbring-Hertz (1988) and Dix<br />

and Webster (1995). A superb film has been<br />

produced by Nordbring-Hertz et al. (1995). Cooke<br />

and Godfrey (1964) have provided a key <strong>to</strong><br />

identification.<br />

25.1.1 Preda<strong>to</strong>ry fungi belonging <strong>to</strong><br />

Ascomycota<br />

Preda<strong>to</strong>ry nema<strong>to</strong>phagous fungi are easy <strong>to</strong><br />

study by means of the sprinkle-plate technique.<br />

A small pinch of soil is added <strong>to</strong> a Petri dish<br />

containing tapwater agar or dilute cornmeal<br />

agar, and the dish is incubated for a few weeks at<br />

room temperature. Saprotrophic nema<strong>to</strong>des<br />

present in the soil will crawl out over the agar<br />

surface, feeding on bacteria. If preda<strong>to</strong>ry fungi<br />

are present in the soil, they develop structures<br />

for trapping nema<strong>to</strong>des, and the trapped dead or<br />

dying animals are easy <strong>to</strong> see with a dissecting<br />

microscope. Conidia will be produced around<br />

colonized nema<strong>to</strong>des, and if these are transferred<br />

<strong>to</strong> standard media such as cornmeal agar,<br />

mycelial colonies will grow. Whilst a few<br />

zygomycetes belonging <strong>to</strong> the order Zoopagales<br />

trap nema<strong>to</strong>des by secreting an adhesive from<br />

undifferentiated hyphae or short hyphal<br />

branches upon contact with nema<strong>to</strong>des<br />

(Drechsler, 1962; Saikawa & Morikawa, 1985),<br />

and two basidiomycete genera also trap nema<strong>to</strong>des<br />

(see Fig. 25.7), the most striking preda<strong>to</strong>ry<br />

species are conidial forms of Ascomycota.<br />

Detailed species descriptions may be found in<br />

the works by van Oorschot (1985) and Rubner<br />

(1996).<br />

The taxonomy of preda<strong>to</strong>ry ascomycetes has<br />

traditionally been based on the morphology of<br />

conidia, especially the number of septa, and the<br />

degree of clustering of conidia on the conidiophore.<br />

However, DNA-based phylogenetic<br />

approaches have confirmed long-held suspicions<br />

that these criteria are artificial, and that more<br />

natural taxa can be obtained by grouping<br />

preda<strong>to</strong>ry ascomycetes according <strong>to</strong> the type of<br />

trap they form. These results have been summarized<br />

and discussed by Scholler et al. (1999). All<br />

known trap-forming ascomycetes described so<br />

far belong <strong>to</strong> the family Orbiliaceae (Hagedorn &<br />

Scholler, 1999), with some species shown <strong>to</strong><br />

produce an apothecium referrable <strong>to</strong> Orbilia<br />

itself (Pfister, 1997). This family is still of<br />

uncertain affinity (incertae sedis) within the<br />

Ascomycota but may belong <strong>to</strong> the Pezizales.<br />

The following trapping structures have been<br />

described.<br />

Adhesive knobs and lateral branches.<br />

Single-celled globose knobs, covered by a sticky<br />

secretion and spaced at intervals along a hypha,<br />

form the morphologically simplest trapping<br />

organs. These knobs are borne directly on the<br />

hypha or on short lateral branches in such a way<br />

that a nema<strong>to</strong>de may become attached <strong>to</strong> several<br />

knobs (Fig. 25.1). A different type of sticky knob<br />

is produced on thin stalks (see Fig. 25.5b).<br />

Sometimes a nema<strong>to</strong>de attached <strong>to</strong> an adhesive<br />

knob may pull it off the subtending hypha by its<br />

violent movement, but respite is only temporary<br />

because penetration of the host may still occur<br />

from such detached knobs. Barron (1977) has<br />

suggested that detachable knobs may provide an<br />

effective means of dispersal. In some species,<br />

lateral knob-bearing branches may develop<br />

in sufficient proximity <strong>to</strong> each other for anas<strong>to</strong>mosis<br />

<strong>to</strong> take place, and this results in the<br />

formation of a primitive two-dimensional sticky<br />

network. Scholler et al. (1999) have proposed<br />

two genera <strong>to</strong> accommodate fungi with<br />

knob-like traps, namely Dactylellina (formerly<br />

Monacrosporium, Arthrobotrys and Dactylella spp.)<br />

for species with stalked knobs, and Gamsylella<br />

(formerly Dactylella spp.) for forms with unstalked<br />

knobs and/or primitive nets.<br />

Adhesive nets. One of the most common<br />

types of trap is a three-dimensional adhesive<br />

network formed by anas<strong>to</strong>mosis of the recurved<br />

hyphal tips of a lateral branch system. The<br />

network is lifted above the general level of the<br />

mycelium. The entire surface of the network is<br />

covered by an adhesive, as shown by scanning<br />

electron microscopy (Fig. 25.2; Nordbring-Hertz,<br />

1972), and a nema<strong>to</strong>de which thrusts its body<br />

in<strong>to</strong> the network is quickly immobilized.<br />

Scholler et al. (1999) have grouped fungi


676 ANAMORPHIC FUNGI<br />

Fig 25.2 Scanning electron micrograph of the adhesive net of<br />

Arthrobotrys oligospora, showing the glue covering the hyphal<br />

surface. Reproduced from Nordbring-Hertz (1972) by<br />

copyright permission of Physiologia Plantarum.Original<br />

pho<strong>to</strong>graph kindly provided by B. Nordbring-Hertz.<br />

Fig 25.1 Gamsylella sp.Hyphae showing short lateral branches<br />

modified as sticky knob traps, with nema<strong>to</strong>des attached at<br />

several points.This species would be called Monacrosporium sp.<br />

in traditional nomenclature because of the single conidium<br />

produced at the end of the conidiophore.<br />

producing traps of this kind in the genus<br />

Arthrobotrys. Adhesive nets are illustrated for<br />

A eudermata (syn. Monacrosporium eudermatum,<br />

Dudding<strong>to</strong>nia flagrans) in Fig. 25.3 and for<br />

A. robusta (Fig. 25.4). Arthrobotrys oligospora<br />

(Fig. 25.2) is by far the most thoroughly investigated<br />

member of the genus.<br />

Non-constricting rings. A number of preda<strong>to</strong>ry<br />

fungi ensnare their prey by three-celled<br />

rings which are formed by recurvature of the tip<br />

of a lateral branch, followed by its anas<strong>to</strong>mosis<br />

with itself. Non-constricting rings have a sticky<br />

inner surface. A nema<strong>to</strong>de thrusting its body<br />

in<strong>to</strong> such a loop may become tightly wedged<br />

inside it, and may find it impossible <strong>to</strong><br />

retract. The point of junction of the loop <strong>to</strong><br />

the subtending hypha is often weak, and the<br />

struggling nema<strong>to</strong>de may detach the loop.<br />

Occasionally, a single nema<strong>to</strong>de bearing several<br />

loops may be seen. The detached loops are still<br />

capable of penetrating and killing the nema<strong>to</strong>de.<br />

The action of this trap is passive, i.e. there is<br />

no inflation of the ring. <strong>Fungi</strong> producing<br />

non-constricting rings also form stalked<br />

adhesive knobs and are included in<br />

Dactylellina (Scholler et al., 1999). An example is<br />

D. hap<strong>to</strong>tyla (syn. Dactylaria candida), shown in<br />

Fig. 25.5.<br />

Constricting rings. The most dramatic type of<br />

trap is the constricting ring trap, which develops


NEMATOPHAGOUS FUNGI<br />

677<br />

Fig 25.3 Arthrobotrys eudermata<br />

(syn. Monacrosporium eudermatum,<br />

Dudding<strong>to</strong>nia flagrans).<br />

(a) Conidiophore with its single<br />

attached conidium. Several<br />

detached conidia are also shown,<br />

one germinating. (b) Trapped<br />

nema<strong>to</strong>de showing infection<br />

bulb and assimilative hyphae.<br />

(c) Adhesive trapping networks.<br />

in the same way as the non-constricting ring but<br />

differs from it in that the three ring cells are able<br />

<strong>to</strong> inflate rapidly following stimulation by<br />

mechanical contact with the inner ring surface.<br />

Inflation occludes the lumen of the ring, severely<br />

constricting any nema<strong>to</strong>de trapped in it. The<br />

surface of constricting rings does not carry any<br />

adhesive. <strong>Fungi</strong> producing constricting rings<br />

have been assigned <strong>to</strong> the genus Drechslerella by<br />

Scholler et al. (1999) and were formerly distributed<br />

in Arthrobotrys and Dactylella. This type of<br />

trap is illustrated in Fig. 25.6.<br />

The mechanism of ring closure is of interest.<br />

The insertion of a fine glass rod in<strong>to</strong> a ring<br />

by means of a micromanipula<strong>to</strong>r, followed by<br />

gentle friction of one of the cells, can trigger off<br />

the closure of the trap. Other stimuli, such as<br />

heat or a stream of dry air, are also effective. The<br />

enlargement of the cells is accompanied by<br />

vacuolation of their contents, and by elastic<br />

stretching of the inside wall of the ring,<br />

whilst the outer wall does not change shape<br />

(Estey & Tzean, 1976). Ring closure is complete<br />

within 0.1 s. Enlargement of the three cells


678 ANAMORPHIC FUNGI<br />

Fig 25.4 (a c) Arthrobotrys robusta.<br />

(a) Conidiophores and conidia.<br />

(b) Mycelium with anas<strong>to</strong>mosing<br />

traps. (c) Nema<strong>to</strong>de caught in trap.<br />

(d) Arthrobotrys oligospora.Two<br />

conidiophores showing the sequence<br />

of conidial development.<br />

making up the ring is not simultaneous, but one<br />

cell inflates a fraction of a second before the<br />

others. By immersing rings in 0.3 0.5 M sucrose<br />

and inducing trap closure by heat, Muller (1958)<br />

succeeded in slowing down the rate of ring<br />

closure by a fac<strong>to</strong>r of 100, so that the process<br />

<strong>to</strong>ok about 10 s. Estimations of the volume<br />

change of the cells during constriction showed<br />

a threefold increase.<br />

Several physiological changes must take place<br />

during closure. Chen et al. (2001) have provided<br />

evidence of a signalling chain in which a physical<br />

stimulus at the inner ring surface activates a<br />

trimeric G protein, and transduction of this<br />

signal via inosi<strong>to</strong>l trisphosphate, Ca 2þ and<br />

calmodulin (see Fig. 12.48) leads <strong>to</strong> the release<br />

of osmotically active molecules within the ring<br />

cells, coupled with the opening of water channels<br />

at the plasma membrane surface. Muller<br />

(1958) estimated that there must be an uptake of<br />

18 000 mm 3 of water in 0.1 s, and water uptake<br />

may therefore take place over much of the


NEMATOPHAGOUS FUNGI<br />

679<br />

Fig 25.5 Dactylellina hap<strong>to</strong>tyla (syn. Dactylaria candida). (a) Conidiophore arising from the substrate, producing a terminal cluster of<br />

septate spindle-shaped conidia. (b) Stalked unicellular knob traps. (c) Non-constricting ring traps. (b,c) <strong>to</strong> same scale.<br />

surface of the ring cells. Since the inner part of<br />

the ring wall rapidly changes shape, it is likely<br />

that there is a slippage of microfibrils making<br />

up the wall. Dowsett et al. (1977) showed that<br />

the inner (luminal) wall of the ring cells in<br />

Drechslerella (Dactylella) brochopaga is thicker than<br />

the outer wall. The luminal wall is initially fourlayered,<br />

but the two external wall layers rupture<br />

during expansion.<br />

The threefold increase in cell volume and the<br />

commensurate increase in surface area, <strong>to</strong>gether<br />

with the process of vacuolation, clearly necessitate<br />

a rapid rearrangement of membrane material<br />

within the cell. Heintz and Pramer (1972)<br />

discovered a labyrinth of membrane-bound<br />

material close <strong>to</strong> the plasma membrane on the<br />

inside of unexpanded ring cells in Drechslerella<br />

(Arthrobotrys) dactyloides. In expanded rings, a<br />

more usual type of plasma membrane organization<br />

was found, suggesting that the membranebound<br />

inclusions had contributed <strong>to</strong> the<br />

formation of the enlarged plasma membrane<br />

(see also Dowsett et al., 1977).<br />

25.1.2 Preda<strong>to</strong>ry fungi belonging <strong>to</strong><br />

Basidiomycota<br />

Nema<strong>to</strong>de-destroying basidiomycetes seem <strong>to</strong><br />

be confined <strong>to</strong> two closely related genera,<br />

Hohenbuehelia and Pleurotus, both belonging <strong>to</strong><br />

the Pleurotaceae in the euagarics clade (p. 541;<br />

Thorn et al., 2000). Both genera contain woodrotting<br />

species forming clamped hyphae.<br />

Hohenbuehelia produces a Nema<strong>to</strong>c<strong>to</strong>nus anamorph<br />

which is of relevance in nema<strong>to</strong>de capture<br />

(Fig. 25.7).<br />

The unusual killing mechanism of Pleurotus<br />

spp. has been described by Barron and<br />

Thorn (1987). Hyphae produce minute unicellular<br />

lollipop-shaped spathulate cells which secrete<br />

droplets of a <strong>to</strong>xin-containing liquid. Various<br />

<strong>to</strong>xins have been identified (p. 682). Nema<strong>to</strong>des<br />

<strong>to</strong>uching such a drop show dramatic symp<strong>to</strong>ms<br />

of a shrinkage of the head region and deformation<br />

of the oesophagus. Onset of symp<strong>to</strong>ms is<br />

within one <strong>to</strong> several minutes and is quickly<br />

followed by lethargy, so that the affected<br />

nema<strong>to</strong>de becomes immobilized close by.<br />

Hyphae of Pleurotus show rapid tropic growth<br />

<strong>to</strong>wards orifices of the paralysed nema<strong>to</strong>de, and<br />

colonization ensues within one <strong>to</strong> several hours<br />

(Fig. 25.7a).<br />

Hohenbuehelia possesses two modes of parasitism.<br />

Hourglass-shaped unicellular mycelial<br />

branches produce a large drop of non-<strong>to</strong>xic glue<br />

by which nema<strong>to</strong>des are trapped in the usual<br />

preda<strong>to</strong>ry way (Fig. 25.7b). In addition, both


680 ANAMORPHIC FUNGI<br />

Fig 25.6 Drechslerella sp.<br />

(a) Conidiophore with a single<br />

terminal conidium typical of the<br />

former genus Monacrosporium.<br />

(b) Three-celled constricting ring<br />

traps. (c) Two traps, one inflated.<br />

(d) Nema<strong>to</strong>de caught in constricting<br />

ring trap. (e) Germinating conidium.<br />

Hohenbuehelia basidiospores and Nema<strong>to</strong>c<strong>to</strong>nustype<br />

conidia produce sticky drops (Barron &<br />

Dierkes, 1977). Drops are formed at the tapering<br />

ends of the conidia which are sometimes bent<br />

at an angle, attaching them <strong>to</strong> the cuticle of<br />

a nema<strong>to</strong>de brushing them (Figs. 25.7c,d).<br />

Hohenbuehelia therefore shows a transition of<br />

preda<strong>to</strong>ry and endoparasitic features (Poloczek<br />

& Webster, 1994). There are also transitions<br />

between Hohenbuehelia and Pleurotus because<br />

Thorn et al. (2000) have described a<br />

Hohenbuehelia sp. producing both sticky knobs<br />

and <strong>to</strong>xin-secreting lollipop-shaped branches.<br />

Unidentified <strong>to</strong>xins have also been shown <strong>to</strong> be<br />

present in the sticky drops on Nema<strong>to</strong>c<strong>to</strong>nus-type<br />

conidia (Giuma et al., 1973).<br />

25.1.3 Endoparasitic nema<strong>to</strong>phagous fungi<br />

Several genera of anamorphic fungi include<br />

forms which are endoparasitic. Although some<br />

of these fungi can be grown in pure culture in<br />

the labora<strong>to</strong>ry, in nature probably most of them


NEMATOPHAGOUS FUNGI<br />

681<br />

Fig 25.7 Nema<strong>to</strong>de-trapping Basidiomycota. (a) Growth of Pleurotus hyphae <strong>to</strong>wards buccal cavities of two nema<strong>to</strong>des killed<br />

at za distance by secreted <strong>to</strong>xins. (b d) Nema<strong>to</strong>c<strong>to</strong>nus state of Hohenbuehelia. (b) Hourglass-shaped cell producing an adhesive.<br />

(c) Production of an awl-shaped conidium from a hypha bearing clamp connections. (d) Conidia attached <strong>to</strong> the cuticle of a dead<br />

nema<strong>to</strong>de by a sticky knob at the tip of a bent neck. Infection has given rise <strong>to</strong> trophic hyphae. (b,c) <strong>to</strong> same scale.<br />

are obligate parasites without a free-living<br />

saprotrophic phase. In some species, the conidia<br />

attach themselves <strong>to</strong> the cuticle of the host and,<br />

on germination, penetrate in<strong>to</strong> the body cavity,<br />

eventually filling it with hyphae. Drechmeria<br />

coniospora (Figs. 25.8a c) has conical conidia<br />

which, at maturity, bear a globose, adhesive<br />

knob at the narrow end. Such conidia readily<br />

adhere <strong>to</strong> the cuticle of a nema<strong>to</strong>de which<br />

brushes past them, and they become preferentially<br />

attached <strong>to</strong> the buccal end of nema<strong>to</strong>des<br />

(Jansson et al., 1985). In some other forms, infection<br />

follows ingestion of a conidium. When the<br />

crescent-shaped conidia of Harposporium anguillulae<br />

are ingested, the pointed end of the spore<br />

becomes lodged in the wall of the oesophagus<br />

and, upon germination, penetrates the body<br />

cavity from within (Aschner & Kohn, 1958). The<br />

conidiophores emerging from the dead host bear<br />

sessile, subglobose phialides (Figs. 25.8d,e). The<br />

mycelium within the host may form chlamydospores<br />

which presumably survive in the soil when<br />

the body of the nema<strong>to</strong>de decays.<br />

Little is known about the taxonomic position<br />

of endoparasitic fungi. Several genera contain<br />

members parasitizing soil invertebrates other<br />

than nema<strong>to</strong>des. For instance, the teleomorph of<br />

Harposporium, Podocrella (syn. Atricordyceps), is<br />

related <strong>to</strong> the insect-parasitic genus Cordyceps in<br />

the Clavicipitaceae (see p. 360). There are also<br />

numerous examples of lower fungi adopting<br />

an endoparasitic mode of life, such as Catenaria<br />

anguillulae (Chytridiomycota) and Myzocytium<br />

spp. (Oomycota) which infect nema<strong>to</strong>des by<br />

means of zoospores, or Hap<strong>to</strong>glossa which<br />

produces the spectacular gun cells shown in<br />

Fig. 3.9.<br />

Although endoparasitic fungi can sometimes<br />

be observed on soil-sprinkle plates, a more<br />

efficient technique for their detection is the<br />

Baermann funnel. A conical funnel is fitted with<br />

rubber tubing at its base, and the opening is<br />

sealed with a clamp. The funnel is lined with<br />

tissue paper or filter paper; the soil sample is<br />

filled in<strong>to</strong> the funnel and submerged in water.<br />

Nema<strong>to</strong>des will migrate through the filter paper


682 ANAMORPHIC FUNGI<br />

Fig 25.8 Endoparasitic<br />

nema<strong>to</strong>phagous fungi. (a c)<br />

Drechmeria coniospora. (a) Dead<br />

nema<strong>to</strong>de containing hyphae, and<br />

bearing emergent conidiophores.<br />

(b) Conidiophores. (c) Mature<br />

conidia showing terminal adhesive<br />

bulbs. (d,e) Harposporium anguillulae.<br />

(d) Dead nema<strong>to</strong>de showing the<br />

extent of internal mycelium, and<br />

external conidiophores.<br />

(e) Enlargement <strong>to</strong> show details<br />

of conidiophores which bear<br />

subglobose phialides and<br />

sickle-shaped phialoconidia.<br />

and congregate at the bot<strong>to</strong>m of the tube,<br />

from which they can be collected on<strong>to</strong> tapwater<br />

agar plates and observed by microscopy.<br />

After a few hours’ incubation, non-motile nema<strong>to</strong>des<br />

may be found, some of them infected with<br />

endoparasitic fungi. These may develop reproductive<br />

structures. Bailey and Gray (1989)<br />

have discussed various isolation techniques for<br />

nema<strong>to</strong>phagous fungi.<br />

25.1.4 Nema<strong>to</strong>de fungus interactions<br />

Stimulation of trap formation<br />

Many nema<strong>to</strong>phagous fungi do not produce<br />

trapping organs in pure culture, but the addition<br />

of nema<strong>to</strong>des or nema<strong>to</strong>de extracts will<br />

induce their development within 24 48 h. The<br />

inducing molecules include small oligopeptides<br />

containing non-polar and aromatic amino<br />

acids, such as the phenylalanyl-valine dipeptide<br />

(Friman et al., 1985). These are present in<br />

exudates from nema<strong>to</strong>des (Nordbring-Hertz,<br />

1977). Other substances, such as horse serum or<br />

yeast extract, are also effective in inducing trap<br />

formation, and in natural substrates such as soil<br />

or dung conidia may germinate directly by<br />

formation of a trap (Persmark & Nordbring-<br />

Hertz, 1997).<br />

In general terms, the readiness <strong>to</strong> form<br />

traps has been found <strong>to</strong> differ between various<br />

genera and is inversely correlated with their<br />

saprotrophic capacity (Cooke, 1964; Barron,<br />

1992). Nordbring-Hertz and Jansson (1984) have<br />

categorized nema<strong>to</strong>phagous fungi in<strong>to</strong> three<br />

groups. Thus, the net-forming genus Arthrobotrys


NEMATOPHAGOUS FUNGI<br />

683<br />

(group 1), credited with the highest competitiveness<br />

in soil, forms its traps only upon induction<br />

by the presence of nema<strong>to</strong>des, whereas producers<br />

of knobs and rings (Dactylellina, Gamsylella<br />

and Drechslerella; group 2) form traps more<br />

readily but have a more limited capacity <strong>to</strong><br />

survive saprotrophically in soil. The obligate<br />

end of the spectrum from saprotrophy <strong>to</strong><br />

parasitism among nema<strong>to</strong>phagous fungi is<br />

occupied by some of the endoparasitic fungi<br />

(group 3). The conidia of these species form sticky<br />

drops constitutively and have no saprotrophic<br />

ability.<br />

Mycelium with traps whether formed<br />

constitutively or induced by the presence of<br />

nema<strong>to</strong>des has been shown <strong>to</strong> attract nema<strong>to</strong>des<br />

significantly more strongly than mycelium<br />

without traps (Field & Webster, 1977; Jansson &<br />

Nordbring-Hertz, 1980). The strength of attraction<br />

can be correlated with the saprotrophic<br />

capacity of the fungus in question, i.e. group 3<br />

fungi exert the highest attraction <strong>to</strong> nema<strong>to</strong>des,<br />

with group 2 fungi being intermediate and<br />

group 1 fungi attracting nema<strong>to</strong>des <strong>to</strong> a lesser<br />

extent even if trap formation has been induced<br />

(Jansson & Nordbring-Hertz, 1979). The identity<br />

of the nema<strong>to</strong>de-attracting substances is as<br />

yet unknown.<br />

Adhesion<br />

The strength of adhesion of a trap or conidium<br />

<strong>to</strong> the nema<strong>to</strong>de prey is remarkable, and much<br />

work has been carried out <strong>to</strong> characterize the<br />

glue involved. This has been summarized by<br />

Nordbring-Hertz (1988) and Tunlid et al. (1992),<br />

who presented evidence of the involvement of<br />

lectins, i.e. proteins that bind <strong>to</strong> specific carbohydrate<br />

residues (recep<strong>to</strong>rs). Different preda<strong>to</strong>ry<br />

and endoparasitic nema<strong>to</strong>phagous fungi differ<br />

in the kind of lectin they produce, and this<br />

may partially account for the specificity of binding<br />

observed in some cases, e.g. in Drechmeria<br />

coniospora whose conidia attach mainly <strong>to</strong> the<br />

buccal end of its prey. It seems that the lectins are<br />

located within the glue on the trap or conidium,<br />

and that the carbohydrate-based recep<strong>to</strong>rs recognized<br />

by the lectins are part of the glycosylation<br />

chains of proteins on the nema<strong>to</strong>de surface.<br />

Toxins<br />

The capture of a nema<strong>to</strong>de by a preda<strong>to</strong>ry fungus<br />

is soon followed by its death, sometimes after<br />

a quick but violent struggle. In the case of<br />

constricting ring traps, the stricture of the body<br />

may well be a contribu<strong>to</strong>ry cause of death,<br />

but there is evidence that <strong>to</strong>xins are also<br />

produced by certain fungi. Stadler et al. (1993)<br />

isolated the common fatty acid linoleic acid as<br />

a nematicidal principle of Arthrobotrys oligospora<br />

and other species. Linoleic acid was produced<br />

at higher amounts by trap-forming cultures<br />

than by uninduced ones, and it was highly <strong>to</strong>xic<br />

against nema<strong>to</strong>des in vitro, even reproducing<br />

the typical symp<strong>to</strong>ms of hyperactivity followed<br />

by paralysis. In the case of Pleurotus spp., two<br />

<strong>to</strong>xins have been isolated, namely trans-2-decenedioic<br />

acid (Kwok et al., 1992) and linoleic acid<br />

(see Anke et al., 1995). The production of antibacterial<br />

antibiotics by numerous nema<strong>to</strong>phagous<br />

fungi has been interpreted as a substrate<br />

defence strategy against bacterial competi<strong>to</strong>rs<br />

during the colonization of a killed nema<strong>to</strong>de<br />

(Anke et al., 1995).<br />

The infection process<br />

Infection of nema<strong>to</strong>des often begins before<br />

the animal is dead. Enzymatic and turgor<br />

pressure-driven mechanisms have been implicated<br />

(Veenhuis et al., 1985; Dijksterhuis et al.,<br />

1990). Since the nema<strong>to</strong>de cuticle consists<br />

mainly of collagen-type proteins and because<br />

serine proteases of the subtilisin type are known<br />

from a wide diversity of nema<strong>to</strong>phagous fungi,<br />

these are generally assumed <strong>to</strong> play an important<br />

role in infection of the host and its<br />

subsequent degradation (Åhman et al., 2002;<br />

Mor<strong>to</strong>n et al., 2004), and their transcription is<br />

enhanced by the presence of nema<strong>to</strong>de cuticle<br />

(Åhman et al., 1996). A subtilisin from A. oligospora<br />

has even been found <strong>to</strong> possess nema<strong>to</strong><strong>to</strong>xic<br />

properties, hinting at several roles which these<br />

proteases may play during infection (Åhman<br />

et al., 2002).<br />

Details of the infection process differ between<br />

preda<strong>to</strong>ry and endoparasitic nema<strong>to</strong>des. In<br />

preda<strong>to</strong>ry species, penetration of the cuticle is<br />

by means of an appressorium or a hypha emitted<br />

by that part of the trap which is in contact


684 ANAMORPHIC FUNGI<br />

with the cuticle. Immediately within the nema<strong>to</strong>de<br />

body, the penetration hypha swells <strong>to</strong><br />

form a globose vesicle, the infection bulb<br />

(see Fig. 25.3b), from which assimilative hyphae<br />

radiate throughout the animal, now dead. The<br />

cy<strong>to</strong>plasm of traps of most preda<strong>to</strong>ry fungi<br />

contains an unusual abundance of dense bodies<br />

(microbodies) identified as peroxisomes. Their<br />

function is currently unknown but may be<br />

related <strong>to</strong> s<strong>to</strong>rage (Veenhuis et al., 1989). Similar<br />

organelles, though of different cy<strong>to</strong>logical<br />

origin, are seen in the assimilative hyphae<br />

within the nema<strong>to</strong>de. It is likely that they<br />

are involved in amino acid assimilation and/or<br />

the degradation of lipids during the digestion<br />

of the nema<strong>to</strong>de contents (Dijksterhuis et al.,<br />

1993).<br />

The sequence of infection-related development<br />

in the endoparasitic conidial species<br />

Drechmeria coniospora and Verticillium balanoides<br />

has been described by Dijksterhuis et al. (1990,<br />

1991) and Sjollema et al. (1993). A hypha growing<br />

out through the adhesive pad at the conidial<br />

apex forms an appressorium on the nema<strong>to</strong>de<br />

cuticle; this mediates penetration. There is no<br />

infection bulb, and the nema<strong>to</strong>de may remain<br />

alive while the trophic hyphae proliferate<br />

within its body. As in preda<strong>to</strong>ry species, numerous<br />

microbodies and lipid droplets are seen<br />

within the trophic hyphae. The extent of mycelium<br />

produced by endoparasitic nema<strong>to</strong>phagous<br />

species in nema<strong>to</strong>des is often limited, with<br />

most of the captured biomass being converted<br />

<strong>to</strong> conidia.<br />

25.1.5 Opportunistic parasites of eggs<br />

and cysts<br />

Numerous saprotrophic soil fungi have been<br />

shown <strong>to</strong> be associated with the eggs and cysts<br />

of nema<strong>to</strong>des, especially sedentary species parasitizing<br />

plant roots (Stiles & Glawe, 1989). Even<br />

though some host specialization may occur, all<br />

fungi colonizing nema<strong>to</strong>de eggs and cysts are<br />

currently considered opportunistic parasites<br />

(Siddiqui & Mahmood, 1996). They reproduce as<br />

conidia and/or chlamydospores, like most soil<br />

fungi. Infection is by means of hyphal tips, and<br />

no specialized infection structures are formed<br />

apart from appressoria in some species.<br />

Nema<strong>to</strong>de eggs contain chitin in addition <strong>to</strong><br />

collagen, and chitinases as well as serine<br />

proteases have been demonstrated in fungal<br />

egg parasites (Mor<strong>to</strong>n et al., 2004). The most<br />

important species as potential biological control<br />

agents are Pochonia chlamydosporia which<br />

produces conidia and multicellular melanized<br />

chlamydospores, and Paecilomyces lilacinus<br />

(Siddiqui & Mahmood, 1996; Kerry & Jaffee,<br />

1997).<br />

25.1.6 Biological control of nema<strong>to</strong>des by<br />

parasitic fungi<br />

Some nema<strong>to</strong>des are serious parasites of plants<br />

and animals, and attempts have been made <strong>to</strong><br />

use all three ecological groups of nema<strong>to</strong>phagous<br />

fungi described above preda<strong>to</strong>ry, endoparasitic<br />

and egg- or cyst-colonizing forms for biological<br />

control. Although the promise held by nema<strong>to</strong>phagous<br />

fungi is high and some success has been<br />

reported, no commercial breakthrough has as<br />

yet been achieved.<br />

Preda<strong>to</strong>ry nema<strong>to</strong>phagous fungi tend <strong>to</strong><br />

show only limited competitiveness in nonnative<br />

soils (Siddiqui & Mahmood, 1996).<br />

Although this may preclude their use in the<br />

biological control of plant-parasitic nema<strong>to</strong>des,<br />

the situation is different with animal parasites.<br />

Many trap-forming fungi occur on dung along<br />

with the larval stages of nema<strong>to</strong>des parasitizing<br />

herbivorous animals. Biological control should<br />

therefore be feasible, but a major obstacle is<br />

the requirement for the spores of such potential<br />

biocontrol agents <strong>to</strong> survive the stringent<br />

passage through the herbivore gut. This<br />

ability has been demonstrated for Arthrobotrys<br />

eudermata (syn. Dudding<strong>to</strong>nia flagrans), which is<br />

unusual among preda<strong>to</strong>ry fungi in forming<br />

thick-walled chlamydospores in addition <strong>to</strong><br />

conidia. There is considerable current interest<br />

in biological control based on feed supplemented<br />

with chlamydospores of A. eudermata<br />

(reviewed by Larsen, 2000).<br />

There are obvious problems in using endoparasitic<br />

nema<strong>to</strong>phagous fungi for biological<br />

control due <strong>to</strong> practical difficulties in producing<br />

sufficient inoculum of many species, and


AQUATIC HYPHOMYCETES (INGOLDIAN FUNGI)<br />

685<br />

the fundamental problem of their low saprotrophic<br />

capacity. The latter feature may be<br />

responsible for observations of a tight coupling<br />

between the population dynamics of root cyst<br />

nema<strong>to</strong>des and those of their fungal endoparasite,<br />

Hirsutella rhossiliensis (Jaffee, 1992).<br />

This species can be grown in pure culture, and<br />

although conidia generated in this way are<br />

not infectious, hyphal inoculum might be<br />

useful for biological control in the future<br />

(Kerry & Jaffee, 1997).<br />

Opportunistic egg and cyst parasites, notably<br />

Pochonia chlamydosporia and Paecilomyces lilacinus,<br />

possess a high ability <strong>to</strong> colonize plant roots in<br />

agricultural soils, and are therefore potentially<br />

useful in the biological control of plant-parasitic<br />

nema<strong>to</strong>des. Both have suppressed root knot<br />

and cyst nema<strong>to</strong>des in experiments under controlled<br />

conditions (Siddiqui & Mahmood, 1996),<br />

although a major problem is that egg masses or<br />

cysts embedded in root tissue cannot be attacked<br />

by either fungus (Kerry, 2000). Pochonia chlamydosporia<br />

has been associated with the suppressive<br />

properties of agricultural soils against the cereal<br />

cyst nema<strong>to</strong>de (Kerry et al., 1982).<br />

25.2 Aquatic hyphomycetes<br />

(Ingoldian fungi)<br />

If a sample of foam from a rapid stream flowing<br />

through deciduous woodland is examined<br />

microscopically, especially in autumn after leaf<br />

fall, it will be found <strong>to</strong> contain a rich variety<br />

of conidia of unusual shape (Fig. 25.9). Many<br />

are quite large, spanning 100 mm or more. These<br />

conidia belong <strong>to</strong> aquatic hyphomycetes which<br />

grow on decaying leaves and twigs. They are<br />

also referred <strong>to</strong> as Ingoldian fungi in honour<br />

of C. T. Ingold who pioneered their study. If<br />

decaying leaves are collected from the stream<br />

and incubated in a shallow layer of water at<br />

a temperature of 10 20°C, numerous conidiophores<br />

of these fungi will develop in a few<br />

hours. Ingoldian fungi have a worldwide<br />

Fig 25.9 Conidia of aquatic<br />

hyphomycetes from river foam.<br />

(a) Volucrispora sp. (b) Ala<strong>to</strong>spora acuminata.<br />

(c) Clava<strong>to</strong>sporalongibrachiata.(d)Tricladium<br />

splendens.(e)Lemonniera aquatica.<br />

(f) Lemonniera terrestris.(g)Articulospora<br />

tetracladia.(h)Clava<strong>to</strong>spora stellata.<br />

(i) Anguillospora crassa.(j)Anguillospora sp.<br />

(k) Heliscus lugdunensis. (l) Unidentified.<br />

(m) Margaritispora aquatica.<br />

(n) Tumularia aquatica.


686 ANAMORPHIC FUNGI<br />

distribution. An excellent guide <strong>to</strong> the group<br />

has been written by Ingold (1975) himself, and<br />

Webster and Descals (1981) have provided keys<br />

<strong>to</strong> the then-known species. Over 300 species<br />

have been described so far. Two common spore<br />

shapes can be recognized the tetraradiate or<br />

branched conidium, and the sigmoid or steephelix<br />

type. Both types of spore may develop<br />

in a variety of different ways, and it is clear<br />

that these spore shapes represent a number of<br />

separate lines of convergent evolution. Evidence<br />

for the view that tetraradiate and sigmoid<br />

propagules have evolved independently several<br />

times is provided by (1) developmental studies,<br />

(2) anamorph teleomorph connections, and<br />

(3) observations of the occurrence of<br />

tetraradiate propagules in unrelated aquatic<br />

organisms.<br />

Table 25.2 lists some of the connections<br />

between aquatic hyphomycetes and different<br />

groups of ascomycetes and basidiomycetes. It is<br />

obvious that some anamorph genera, e.g.<br />

Tricladium or Anguillospora, are unnatural (polyphyletic),<br />

i.e. they are not made up of related<br />

species. The teleomorphs for which connections<br />

have been established develop more readily on<br />

submerged or partially submerged wood than on<br />

leaves, although a few leaf-borne teleomorphic<br />

states have been described. The pleomorphic<br />

nature of Ingoldian fungi, as in other fungal<br />

groups, presents problems of nomenclature.<br />

Strictly, the name adopted should be that of<br />

Table 25.2. Examples of the taxonomic diversity of Ingoldian fungi (based on Webster, 1992 and Descals<br />

et al., 1998).<br />

Anamorph genus Teleomorph genus Taxonomic affinity<br />

Branched conidial Ascomycota<br />

Actinosporella Miladina Pezizales<br />

Anavirga Vibrissea Helotiales<br />

Articulospora Hymenoscyphus Helotiales<br />

Casaresia Mollisia Helotiales<br />

Clavariopsis Massarina Dothideales (Loculoascomycetes)<br />

Dwayaangam Orbilia Orbiliaceae<br />

Geniculospora Hymenoscyphus Helotiales<br />

Tricladium<br />

Hymenoscyphus,<br />

Helotiales<br />

Hydrocina,Cudoniella<br />

Varicosporium Hymenoscyphus Helotiales<br />

Branched conidial Basidiomycota<br />

Taeniomyces Fibulomyces Polyporoid clade (Homobasidiomycetes)<br />

Ingoldiella Sis<strong>to</strong>trema Polyporoid clade (Homobasidiomycetes)<br />

Crucella Camp<strong>to</strong>basidium Atractiellales (Urediniomycetes)<br />

Sigmoid conidial Ascomycota<br />

Anguillospora<br />

Mollisia, Pezoloma,<br />

Helotiales<br />

Hymenoscyphus, Loramyces<br />

Anguillospora Orbilia Orbiliaceae<br />

Anguillospora Massarina Dothideales (Loculoascomycetes)<br />

Flagellospora Nectria Hypocreales (Pyrenomycetes)<br />

Conidial Ascomycota of other shapes<br />

Heliscus,Cylindrocarpon Nectria Hypocreales (Pyrenomycetes)<br />

Dimorphospora Hymenoscyphus Helotiales<br />

Tumularia Massarina Dothideales (Loculoascomycetes)


AQUATIC HYPHOMYCETES (INGOLDIAN FUNGI)<br />

687<br />

the teleomorph, but in practice, because the<br />

teleomorphs are less frequently encountered<br />

than the anamorphs, it is the name of the<br />

latter which is generally used.<br />

25.2.1 Tetraradiate and other branched<br />

conidia<br />

A few examples of the development of branched<br />

conidia illustrate great variation which can be<br />

followed by making observations of spore development<br />

over a period of a few hours on leaf<br />

fragments bearing conidiophores or from pieces<br />

of agar culture incubated in water. In some cases<br />

development is aided by placing culture pieces<br />

in special flow cells which permit microscopic<br />

observations <strong>to</strong> be made under continuous water<br />

flow over time.<br />

Phialidic tetraradiate conidia<br />

A good example is Lemonniera aquatica<br />

(Fig. 25.10a), probably the most common of the<br />

six species in this genus. Conidiophores develop<br />

from mycelium embedded in the leaf tissues or<br />

from chlamydospores or sclerotia, and terminate<br />

in 1 3 phialides. From the tip of the phialide<br />

a tetrahedral conidium primordium develops,<br />

and the four corners of the tetrahedron extend<br />

simultaneously <strong>to</strong> form cylindrical arms which<br />

may become septate. The mature conidium is<br />

thus attached centrally <strong>to</strong> the phialide at the<br />

point of divergence of the arms. When the firstformed<br />

conidium is detached <strong>to</strong> be carried away<br />

by water currents, a second conidium develops,<br />

and others follow.<br />

Phialides of Ala<strong>to</strong>spora (Fig. 25.10b) develop<br />

singly at the tips of short, inconspicuous conidiophores.<br />

Midway along the length of an<br />

elongated spore initial, two divergent lateral<br />

arms arise and extend simultaneously.<br />

Heliscus lugdunensis (Fig. 25.10c) is a common<br />

early colonizer of the bark of twigs which have<br />

fallen in<strong>to</strong> streams. Its conidiophores develop on<br />

sporodochium-like pustules and branch repeatedly,<br />

terminating in phialides. Conidia which<br />

develop underwater are clove-shaped with short,<br />

conical projections at the upper end, whereas<br />

more cylindrical conidia are produced under<br />

aerial conditions, e.g. on a twig incubated in<br />

a moist chamber or on agar culture. The<br />

Fig 25.10 Three phialidic aquatic hyphomycetes.<br />

(a) Lemonniera aquatica.(b)Ala<strong>to</strong>spora acuminata.(c)Heliscus<br />

lugdunensis.<br />

teleomorph is Nectria lugdunensis which forms<br />

bright red perithecia on half-submerged twigs.<br />

Blastic tetraradiate conidia<br />

There are numerous examples of blastic conidial<br />

development in aquatic hyphomycetes.<br />

Articulospora tetracladia (Fig. 25.11a) forms short<br />

conidiophores extending from mycelium within<br />

a leaf. At the tip of the conidiophore the first<br />

arm develops as a cylindrical bud. At the apex<br />

of this first arm, three further cylindrical buds<br />

develop in turn. A narrow constriction or<br />

joint marks the point of attachment of these<br />

later-formed arms <strong>to</strong> the first (hence the name<br />

Articulospora, a jointed spore). The mycelium and


688 ANAMORPHIC FUNGI<br />

Fig 25.11 Two blastic aquatic hyphomycetes. (a) Articulospora<br />

tetracladia.The arms of the conidia develop successively.<br />

(b) Clavariopsis aquatica.The <strong>to</strong>p-shaped body of the conidium<br />

develops first, followed by simultaneous development of the<br />

three thinner arms.<br />

Fig 25.12 Tricladium splendens. (a) Stages in blastic<br />

development of conidia. A club-shaped main axis develops<br />

lateral arms successively from different points along its length.<br />

(b) Mature detached conidia.<br />

conidia of Articulospora are hyaline. The teleomorph<br />

is an inoperculate discomycete,<br />

Hymenoscyphus tetracladius.<br />

Clavariopsis aquatica (Fig. 25.11b) has dark<br />

mycelium and conidia. The conidia have a<br />

broad, obconical body with a rounded tip<br />

bearing three cylindrical arms which develop<br />

simultaneously. The mature conidium usually<br />

has a single septum in the central body. In<br />

culture, a spermogonial state has been found,<br />

a dark-coloured pycnidium containing minute,<br />

colourless spermatia. The teleomorph belongs <strong>to</strong><br />

the genus Massarina (Loculoascomycetes).<br />

Tricladium splendens (Fig. 25.12) also has darkcoloured<br />

mycelium and conidia. The apex of the<br />

conidiophore develops a club-shaped swelling<br />

which becomes septate and forms the main axis<br />

of the conidium. A bud develops at one point on<br />

the main axis, <strong>to</strong> be followed by a second, at<br />

a different point. The arms taper and are<br />

constricted where they join the main axis. Its<br />

teleomorph is Hymenoscyphus splendens.<br />

Tetrachaetum elegans (Fig. 25.13) has a hyaline<br />

mycelium and conidia which are relatively large,<br />

spanning up <strong>to</strong> 200 mm. The conidium develops<br />

by the curvature of the main axis which is<br />

narrowly cylindrical. Two laterals arise at a<br />

common point about halfway along the main<br />

axis, and develop simultaneously. Varicosporium<br />

elodeae (Fig. 25.14) and Dendrospora erecta<br />

(Fig. 25.15) bear blas<strong>to</strong>conidia which are more<br />

highly branched, with further branches developing<br />

from the primary laterals.<br />

Branched conidia with clamp connections and<br />

dolipore septa<br />

A number of branched conidia found in water or<br />

foam have clamp connections at their septa,<br />

showing that they are basidiomycetes. They


AQUATIC HYPHOMYCETES (INGOLDIAN FUNGI)<br />

689<br />

Fig 25.13 Tetrachaetum elegans.The main axis of the conidium<br />

bends and, at the point of curvature, two lateral arms arise<br />

simultaneously.<br />

obtained in cultures derived from a conidium.<br />

The teleomorph is a species of Fibulomyces<br />

probably belonging <strong>to</strong> the polyporoid clade.<br />

Ingoldiella hamata (Fig. 25.17), a tropical<br />

aquatic fungus, has large, dikaryotic conidia<br />

with numerous clamp connections. The basidial<br />

state is Sis<strong>to</strong>trema hamatum (polyporoid clade)<br />

with eight-spored basidia. Single basidiospores<br />

germinate <strong>to</strong> form monokaryotic mycelia on<br />

which monokaryotic conidia develop. These<br />

closely resemble dikaryotic conidia but lack<br />

clamp connections.<br />

There are other aquatic hyphomycetes with<br />

basidiomyce<strong>to</strong>us affinities indicated by the<br />

possession of dolipore septa within the mycelium<br />

and the conidium. Examples include<br />

Dendrosporomyces prolifer and D. splendens which<br />

have conidia with a strong morphological resemblance<br />

<strong>to</strong> Dendrospora, and Tricladiomyces malaysianum<br />

with spores resembling those of a<br />

Tricladium but with dolipore septa (Nawawi,<br />

1985).<br />

25.2.2 Sigmoid conidia<br />

A similar range of different types of conidial<br />

on<strong>to</strong>geny can be demonstrated for sigmoid<br />

conidia.<br />

Phialoconidia<br />

Flagellospora curvula (Fig. 25.18a) has narrow,<br />

sigmoid phialoconidia developing from phialides<br />

on a sparsely branched conidiophore. A more<br />

richly branched, penicillate arrangement of<br />

phialides is found in F. penicillioides (Fig. 25.18b),<br />

which has a Nectria teleomorph.<br />

Fig 25.14 Varicosporium elodeae. Branched<br />

blas<strong>to</strong>conidia formedby repeatedbranching of the lateral arms<br />

which develop mostly from one side of the main axis.<br />

include Taeniomyces (Fig. 25.16) which has a<br />

conidium somewhat resembling a Tricladium,<br />

but with a single clamp connection at the<br />

septum lying between the two arms. The conidium<br />

is dikaryotic, and the basidial state has been<br />

Blas<strong>to</strong>conidia<br />

Anguillospora has blastic, sigmoid conidia, but<br />

evidence from the known teleomorphs indicates<br />

that this form-genus is very heterogeneous,<br />

including species from unrelated groups of<br />

Ascomycota (see Table 25.2). There are also<br />

differences in the mechanism of conidial<br />

separation. Anguillospora longissima (Fig. 25.19)<br />

has dark mycelium and conidia. The conidia<br />

develop as club-shaped swellings from the<br />

apices of conidiophores. The conidium becomes<br />

septate and helically curved. Conidial separation<br />

is rhexolytic, brought about by the collapse


690 ANAMORPHIC FUNGI<br />

Fig 25.15 Dendrospora erecta.<br />

Much-branched blas<strong>to</strong>conidia formed<br />

by repeated branching of lateral arms<br />

which develop on all sides of the main<br />

axis of the conidium.<br />

Fig 25.16 Taeniomyces gracilis.<br />

(a) Mature detached conidia. Note<br />

the clamp connection between the<br />

lateral arms. (b) Basidial state<br />

(Fibulomyces sp.).<br />

of a special separating cell at the base of the<br />

conidium. The contents of the separating cell<br />

disintegrate, and the cell wall breaks down at<br />

a line of weakness near the middle. When the<br />

conidium separates, it carries at its base a little<br />

collar which represents half of the empty separating<br />

cell. Similarly, the apex of the conidiophore<br />

bears a collar after detachment of the<br />

first conidium. The conidiophore may develop a<br />

second conidium by percurrent extension<br />

through the remnants of the first separating cell<br />

and, after several conidia have been formed, a<br />

succession of collars may be found at the tip of<br />

the conidiophore. The teleomorph of A. longissima<br />

is a pseudothecium-forming species of<br />

Massarina found on twigs in streams.<br />

A second species, A. furtiva, has conidia which<br />

closely resemble those of A. longissima, but cultures<br />

derived from such conidia have developed<br />

apothecia of an inoperculate discomycete,<br />

Pezoloma sp. In A. furtiva there is no separating<br />

cell and conidium separation is schizolytic,<br />

i.e. by dissolution of a septum at its base<br />

(Descals et al., 1998). In A. crassa, which has fatter


AQUATIC HYPHOMYCETES (INGOLDIAN FUNGI)<br />

691<br />

Fig 25.17 Ingoldiella hamata. (a) Developing dikaryotic conidia.<br />

Note that the septa bear clamp connections. (b) Mature<br />

dikaryotic conidia. (c) Monokaryotic conidium lacking clamp<br />

connections.<br />

conidia, separation is also brought about by<br />

septum dissolution, and this species also has an<br />

inoperculate discomycete as its teleomorph,<br />

Mollisia uda.<br />

Lunulospora curvula (Fig. 25.18c) has crescentshaped<br />

blas<strong>to</strong>conidia which develop from<br />

specialized conidiogenous cells at the apex of<br />

dark conidiophores. This fungus is more<br />

common in warmer countries than in temperate<br />

regions, and in Britain its season of maximum<br />

abundance is in late summer and autumn.<br />

25.2.3 Other types of spore<br />

Not all aquatic hyphomycetes have branched<br />

or sigmoid conidia. Margaritispora aquatica<br />

forms hyaline, globose phialoconidia bearing<br />

a few conical protrusions (Fig. 25.20a), whilst


692 ANAMORPHIC FUNGI<br />

Fig 25.18 Three aquatic<br />

hyphomycetes with sigmoid conidia.<br />

(a) Flagellospora curvuula<br />

conidiophores with phialides and<br />

sigmoid phialoconidia. (b) Flagellospora<br />

penicillioides conidiophores with<br />

phialides and phialoconidia.<br />

(c) Lunulosporacurvula,showingblastic<br />

development of crescent-shaped<br />

conidia.<br />

Fig 25.19 Anguillospora longissima,anaquatic<br />

hyphomycete with blastic sigmoid conidia.<br />

(a) Detachment of conidium showing the<br />

remnants of the separating cell, and percurrent<br />

proliferation. (b) Developing conidia.The arrow<br />

marks a separating cell. (c) Mature detached<br />

conidia.


AQUATIC HYPHOMYCETES (INGOLDIAN FUNGI)<br />

693<br />

Tumularia aquatica (Fig. 25.20b) has pear-shaped<br />

or broadly fusiform blas<strong>to</strong>conidia which separate<br />

by septal dissolution. The teleomorph of this<br />

fungus is Massarina aquatica which forms pseudothecia<br />

on submerged wood in streams.<br />

25.2.4 The significance of spore shape<br />

in aquatic fungi<br />

We have seen that tetraradiate conidia may<br />

develop in a variety of ways and are produced<br />

both by Ascomycota and Basidiomycota.<br />

Tetraradiate propagules have also been found<br />

as secondary conidia of En<strong>to</strong>mophthoraceae<br />

attacking aquatic insects (see Fig. 7.42; Descals<br />

et al., 1981), and as basidiospores in the marine<br />

fungi Digita<strong>to</strong>spora marina (Homobasidiomycetes)<br />

and Nia vibrissa (gasteromycetes) (Doguet, 1962,<br />

1967). The brown alga Sphacelaria also forms<br />

tetraradiate propagules. There is thus ample<br />

evidence for the view that this type of structure<br />

has evolved repeatedly in aquatic environments.<br />

The functional significance of convergent evolution<br />

of tetraradiate and sigmoid spore shapes has<br />

been discussed by Webster (1987). Experimental<br />

studies have shown that tetraradiate propagules<br />

are more effectively trapped by impaction on<strong>to</strong><br />

underwater objects than spores of more conventional<br />

shape. This is because a tetraradiate<br />

propagule making contact with a surface will<br />

achieve a three-point landing, a very stable form<br />

of attachment. Attachment is aided by secretion<br />

of mucilage at the tips of the arms which quickly<br />

develop ‘appressoria’ and germ tubes, but the<br />

fourth arm not in contact with a surface fails<br />

<strong>to</strong> differentiate in this way (Read et al., 1991,<br />

1992a,b; Jones, 1994). Once appressoria have<br />

been formed, the propagules are very resistant<br />

<strong>to</strong> detachment.<br />

Trapping efficiency may not be the only<br />

advantage of a tetraradiate propagule because<br />

these spores may also be found among leaves in<br />

terrestrial habitats (Bandoni, 1972; Park, 1974).<br />

Fig 25.20 (a) Margaritispora aquatica,<br />

conidiophores, phialides and<br />

phialoconidia. (b) Tumularia aquatica,<br />

conidiophores and blas<strong>to</strong>conidia.


694 ANAMORPHIC FUNGI<br />

Bandoni (1974) has advanced the idea that<br />

tetraradiate spores may be adapted <strong>to</strong> movement<br />

in surface films of water.<br />

Although sigmoid spores are less efficiently<br />

trapped than tetraradiate spores, they, <strong>to</strong>o, can<br />

develop in different ways and in unrelated<br />

groups of fungi which have adopted an aquatic<br />

habit, suggesting that their shape has selective<br />

value. Observations on sigmoid spores moving<br />

with the current flow in flat capillary tubes show<br />

that as they approach a surface they tumble end<br />

over end and come <strong>to</strong> rest with one spore tip in<br />

contact with the surface. Immediately after<br />

arrest, the spore swings parallel <strong>to</strong> the current,<br />

thus minimizing the shear forces acting <strong>to</strong><br />

detach the spore (Webster & Davey, 1984). As in<br />

tetraradiate spores, the tips of the arms of<br />

sigmoid spores secrete mucilage, possibly in<br />

response <strong>to</strong> a thigmotropic stimulus associated<br />

with their tumbling movements. Mucilage is also<br />

present on the outside of the sigmoid spore of<br />

Mycocentrospora filiformis prior <strong>to</strong> it making<br />

contact with a surface (Au et al., 1996). Current<br />

flow forces the spore in<strong>to</strong> contact with the<br />

surface at a second point along its length.<br />

Germination occurs by the development of<br />

a germ tube from that end of the spore which<br />

is in contact with the surface. Thus, in contrast<br />

<strong>to</strong> the three-point contact associated with tetraradiate<br />

spore shape, sigmoid spores make two<br />

points of contact with surfaces <strong>to</strong> which they<br />

adhere.<br />

25.2.5 Spores in stream foam<br />

Foam is an effective trap for both tetraradiate<br />

and sigmoid spores (see Fig. 25.9). Spores<br />

suspended in water or caught in stream foam<br />

rarely germinate and many studies on the<br />

distribution and seasonal abundance of aquatic<br />

hyphomycetes using preserved foam samples<br />

have been made. In experiments in which air<br />

bubbles were passed through concentrated<br />

suspensions of conidia, the concentration of<br />

suspended spores fell very rapidly. Tetraradiate<br />

conidia were removed more readily than conidia<br />

of sigmoid or other shape (Iqbal & Webster,<br />

1973a). Comparisons of spores collected in foam<br />

or by Millipore filtration from the same stream<br />

indicates that the spore content of foam overrepresents<br />

the tetraradiate type of conidium<br />

in relation <strong>to</strong> other spore types known <strong>to</strong> be<br />

present. It cannot be assumed that all propagules<br />

found in stream foam originate from within<br />

the stream. Some come from fungi growing on<br />

the living leaves of riparian trees and the spores<br />

are brought in<strong>to</strong> the stream in raindrops or<br />

from rainwater draining down the tree trunks.<br />

Such fungi have been distinguished as terrestrial<br />

aquatic hyphomycetes (Ando & Tubaki,<br />

1984a,b).<br />

25.2.6 Adaptations of Ingoldian fungi <strong>to</strong><br />

the aquatic habitat<br />

Ingoldian fungi represent an ecological group of<br />

fungi sharing a common habitat, typically leaves<br />

and twigs in rapidly flowing streams. It is<br />

believed that they have been derived from<br />

terrestrial ances<strong>to</strong>rs and are able <strong>to</strong> colonize<br />

their habitat by virtue of a number of adaptations.<br />

These include effective spore attachment<br />

mechanisms associated with tetraradiate and<br />

sigmoid spore shape, rapid germination, and<br />

rapid growth and sporulation enhanced by<br />

turbulence and rapid water flow (Webster &<br />

Towfik, 1972; Sanders & Webster, 1980).<br />

Physiological adaptations include the possession<br />

of a range of enzymes enabling them <strong>to</strong> degrade<br />

their substrata (see below), and an ability <strong>to</strong><br />

grow at low temperatures, sometimes approaching<br />

0°C, so that they can continue growth<br />

and sporulation on submerged deciduous tree<br />

leaves after the autumn pulse of leaf fall<br />

(Koske & Duncan, 1974). Despite their typical<br />

environment, many aquatic hyphomycetes can<br />

survive for several weeks on dried leaves<br />

previously colonized in streams and brought<br />

out by flooding or by falling water levels (Sanders<br />

& Webster, 1978). It is possible that overland<br />

dispersal is achieved if colonized dried leaves<br />

are blown about by wind and re-deposited in<br />

streams.<br />

25.2.7 Ecophysiological studies<br />

There have been extensive studies on the ecology<br />

and physiology of Ingoldian hyphomycetes<br />

stimulated by the discovery that they play an


AQUATIC HYPHOMYCETES (INGOLDIAN FUNGI)<br />

695<br />

essential role in rendering leaves which they<br />

colonize in streams more palatable and nutritious<br />

<strong>to</strong> aquatic invertebrates feeding on them<br />

(Bärlocher, 1992; Dix & Webster, 1995; see below).<br />

Ecological studies have been facilitated by the<br />

fact that these fungi can, in general, be recognized<br />

<strong>to</strong> a high degree of certainty from their<br />

spores, which is not the case for most other<br />

groups of fungi.<br />

Distribution<br />

Ingoldian hyphomycetes have a worldwide distribution<br />

from the equa<strong>to</strong>r <strong>to</strong> the Arctic. They are<br />

found most frequently in babbling brooks overhung<br />

by deciduous trees and are less abundant<br />

in wider rivers or where streams flow through<br />

afforested regions in which trees have been clearfelled<br />

on both sides (Metwalli & Shearer, 1989).<br />

They are relatively infrequent in streams in<br />

mountainous or moorland areas devoid of<br />

riparian trees, but can grow there on plants<br />

such as rushes (Juncus spp.). As rivers flow<br />

<strong>to</strong>wards the sea and the brackish condition is<br />

encountered in the estuaries, Ingoldian hyphomycetes<br />

decline in frequency. A few species grow<br />

in lakes, but the lotic (flowing) habitat is<br />

preferred <strong>to</strong> the lentic (smooth).<br />

Substrates<br />

The main substrates for aquatic hyphomycetes<br />

are leaves of deciduous trees. In general these<br />

fungi show little host specificity but certain tree<br />

leaves support a richer fungal population with<br />

more abundant sporulation than others (Gulis,<br />

2001). A particularly rich mycota is associated<br />

with the leaves of alder (Alnus glutinosa), a<br />

common riparian tree, and this is possibly correlated<br />

with their relatively high nitrogen content<br />

due <strong>to</strong> the fact that Alnus can fix gaseous N 2 .<br />

In contrast, the leaves of beech (Fagus sylvatica)<br />

are a poor substrate. Needles of conifers are<br />

resistant <strong>to</strong> colonization by aquatic hyphomycetes,<br />

related partly <strong>to</strong> their thick cuticles and<br />

also <strong>to</strong> the presence of compounds inhibi<strong>to</strong>ry <strong>to</strong><br />

mycelial growth (Bärlocher & Oertli, 1978a,b).<br />

Wood which has fallen in<strong>to</strong> streams is an<br />

important substratum because it is more enduring<br />

than leaves, which decompose more rapidly<br />

and are consumed by animals or scoured by<br />

water currents, so that they may not survive in<br />

great quantity <strong>to</strong> provide inoculum for the next<br />

season’s autumnal input. Additionally, colonized<br />

wood is a substratum for the development of<br />

teleomorphs (Shearer, 1992). Living riparian tree<br />

roots also support a population of aquatic<br />

hyphomycetes (Fisher et al., 1991; Sridhar &<br />

Bärlocher, 1992a,b), with roots of alder and<br />

willow (Salix) particularly important because<br />

they extend in<strong>to</strong> streams. Aquatic hyphomycetes<br />

have even been reported from beech roots<br />

growing in woodland soil (Waid, 1954). In<br />

lowland streams and rivers a few species may<br />

colonize living and decaying leaves of in-stream<br />

macrophytes (Kirby et al., 1990). Treeholes,<br />

cavities formed within tree trunks where the<br />

stumps of fallen branches have decayed, intermittently<br />

fill with rain water. Samples of the<br />

water and the leaf and other debris which<br />

accumulates in treeholes reveal the frequent<br />

presence of Ingoldian fungi such as Ala<strong>to</strong>spora<br />

acuminata (Gönczöl & Révay, 2003).<br />

Spore concentrations in streams<br />

Concentrations of aquatic hyphomycete spores<br />

in streams can be readily estimated by the<br />

filtration of water samples through a Millipore<br />

filter (preferably with 8 mm pore size, <strong>to</strong> facilitate<br />

rapid filtration) followed by treatment which<br />

stains the spores and renders the filter transparent.<br />

Concentrations reach a peak soon after the<br />

main period of autumnal deciduous leaf fall<br />

in temperate countries. Spore counts as high as<br />

2 3 l0 4 l 1 have been made in Oc<strong>to</strong>ber and<br />

November in Britain and elsewhere (Iqbal &<br />

Webster, 1973b). As the leaves are decomposed,<br />

consumed or swept downstream, the concentration<br />

of spores in suspension may fall <strong>to</strong> undetectable<br />

levels. This points <strong>to</strong> the significance of<br />

colonization of more enduring woody substrata<br />

and also growth in the roots of riparian trees<br />

which extend in<strong>to</strong> the water.<br />

Feeding of invertebrates on aquatic<br />

hyphomycetes<br />

It is now becoming clear that aquatic hyphomycetes<br />

play an important role in the cycling of<br />

nutrients in streams (e.g. Kaushik & Hynes, 1968,<br />

1971; Bärlocher & Kendrick, 1973a,b; Suberkropp


696 ANAMORPHIC FUNGI<br />

& Klug, 1976, 1980). Streams bordered by trees<br />

receive the bulk of their fixed carbon input not<br />

from attached macrophytes or algae, but from<br />

leaves, twigs and other debris shed by trees and<br />

other plants. Such material is relatively poor in<br />

nitrogen or is otherwise unpalatable <strong>to</strong> the<br />

invertebrate animal population. The colonization<br />

of leaf litter by aquatic fungi and bacteria is<br />

an important part of the ‘processing’ which<br />

makes it a more attractive substrate <strong>to</strong> animals<br />

(Berrie, 1975). There are two main reasons for<br />

this, namely pre-digestion and enrichment in<br />

organic nitrogen.<br />

First, the activities of the fungi soften the leaf<br />

tissues. Aquatic hyphomycetes possess a range of<br />

pec<strong>to</strong>lytic, cellulolytic, proteolytic and ligninolytic<br />

enzymes capable of degrading leaf tissues<br />

(Suberkropp & Klug, 1980; Chamier, 1985; Zemek<br />

et al., 1985; Gessner et al., 1997). Colonized<br />

softened leaves are more easily grazed and<br />

shredded than uncolonized leaves by aquatic<br />

invertebrates such as Gammarus and Asellus and<br />

by the larvae of aquatic insects feeding directly<br />

on leaf tissue or on particulate organic matter<br />

released in<strong>to</strong> streams as leaves decay. In general,<br />

aquatic animals do not possess enzymes capable<br />

of degrading the cell walls of leaf tissue, but the<br />

enzymes present within ingested leaf fragments<br />

may continue <strong>to</strong> be active within the animals’<br />

guts. When presented with a choice of uncolonized<br />

or colonized leaf tissue, either as separate<br />

discs or as patches on the same leaf, caddis fly<br />

larvae feed preferentially on colonized tissue<br />

(Arsuffi & Suberkropp, 1985).<br />

Second, the protein content of leaf material is<br />

enhanced by microbial colonization. Aquatic<br />

fungi can concentrate inorganic nitrogen<br />

present in solution in the water at low concentrations<br />

but in large <strong>to</strong>tal amounts and, making<br />

use of the organic matter in the leaves, manufacture<br />

microbial protein. <strong>Fungi</strong> can make up<br />

over 90% of the microbial biomass which develops<br />

on decomposing leaves in streams. Aquatic<br />

animals may feed directly on the fungal mycelium<br />

and on fungal spores. Detritivorous animals<br />

such as Gammarus pulex and Asellus aquaticus<br />

fed on a fungus diet make a much greater<br />

weight increase than those fed solely on a diet<br />

of uncolonized leaves. They are also more<br />

fecund, i.e. produce more eggs per brood<br />

(Graca et al., 1993). In order <strong>to</strong> sustain their<br />

restricted growth rate on leaf diets, the animals<br />

consume about 10 times more leaf material<br />

(by dry weight) than individuals fed on fungus<br />

diets. Thus aquatic hyphomycetes play a very<br />

important role as intermediaries in the diet<br />

of aquatic invertebrates and their major food<br />

source, the leaves of riparian trees. Since aquatic<br />

invertebrates in turn provide the food source<br />

of other animals, including fish, the activities<br />

of aquatic hyphomycetes are vital <strong>to</strong> the food<br />

chain in maintaining stream productivity.<br />

25.3 Aero-aquatic fungi<br />

If leaves and twigs from the mud surface of<br />

stagnant pools or slow-running ditches are<br />

rinsed and incubated at room temperature in<br />

a humid environment (e.g. a Petri dish or plastic<br />

box lined with wet blotting paper), fungi with<br />

very characteristic large conidia usually develop<br />

within a few days. The common feature of the<br />

conidia of these fungi is that they trap air as they<br />

develop, which assists in floating off the conidia<br />

if the substratum is submerged in water. Conidia<br />

of this type have been termed bubble-trap<br />

propagules (Michaelides & Kendrick, 1982).<br />

Such fungi grow vegetatively on leaves and<br />

twigs, often in water with quite low amounts<br />

of dissolved oxygen. Under submerged conditions<br />

these fungi do not sporulate, but do so<br />

only after incubation under aerial conditions<br />

in which a moist interface between air and water<br />

is provided, as might happen at a pond margin<br />

as the water dries up and previously submerged<br />

twigs or leaves become exposed <strong>to</strong> air. They have<br />

therefore been termed aero-aquatic fungi.<br />

Aero-aquatic fungi are an ecological group<br />

of organisms without phylogenetic coherence,<br />

as shown in Table 25.3. Although the taxonomy<br />

of the group is fairly well known (see e.g. Linder,<br />

1929; Moore, 1955; Webster & Descals, 1981;<br />

Goos, 1987; Voglmayr, 2000), it is likely that<br />

many more species remain <strong>to</strong> be discovered.<br />

Careful studies should also reveal new ascomyce<strong>to</strong>us<br />

teleomorphs because several species


AERO-AQUATIC FUNGI<br />

697<br />

Table 25.3. Examples of the taxonomic diversity of aero-aquatic fungi.<br />

Anamorph genus Teleomorph genus Taxonomic affinity<br />

Oomycota<br />

Medusoides (oogonium) Pythioge<strong>to</strong>naceae (Pythiales)<br />

Ascomycota<br />

Clathrosphaerina Hyaloscypha Helotiales<br />

Helicoon Orbilia Orbiliaceae<br />

Helicodendron Mollisia, Hymenoscyphus Helotiales<br />

Helicodendron Lambertella, Herpotrichiella Chae<strong>to</strong>thyriales (see p.484)<br />

Helicodendron Tyrannosorus Dothideales (Loculoascomycetes)<br />

Pseudaegerita Hyaloscypha Helotiales<br />

Basidiomycota<br />

Aegerita Bulbillomyces Polyporoid clade<br />

Aegeritina Subulicystidium Polyporoid clade<br />

(Unknown) Limnoperdon (basidiocarp) Gasteromycetes<br />

Akenomyces (Sclerotium) (Unknown) Incertae sedis<br />

of aero-aquatic fungi form microconidial synanamorphs<br />

in culture, the spores of which fail<br />

<strong>to</strong> germinate. These are probably spermatia.<br />

25.3.1 Development of propagules<br />

There are various ways in which conidia of aeroaquatic<br />

hyphomycetes may develop. In Helicoon<br />

(Fig. 25.21a) they develop as cylindrical or barrelshaped<br />

spirals. The conidia vary in colour<br />

from hyaline <strong>to</strong> black. The direction of coiling<br />

of the spirals (looking upwards from the apex of<br />

the conidiophore) is clockwise in H. richonis<br />

whereas in some other helicosporous fungi the<br />

direction of coiling is counter-clockwise. The<br />

direction appears <strong>to</strong> be constant for a given<br />

species. In Helicoon the conidia themselves do<br />

not branch, but in Helicodendron (Hd.) which is<br />

clearly a polyphyletic anamorph genus, the<br />

conidia may bear further conidia as lateral<br />

branches (Figs. 25.21b,c; 25.22a). Beverwijkella<br />

pulmonaria, probably an anamorphic ascomycete<br />

(Fig. 25.23a), forms uni- or bi-lobed, balloon-like<br />

structures. At the surface, aggregates of dark,<br />

thick-walled, tightly packed cells develop, with<br />

air trapped inside the cavity which they enclose<br />

(Michaelides & Kendrick, 1982). Spirosphaera,<br />

another anamorphic ascomycete, achieves the<br />

same end by the formation of globose<br />

propagules made up of richly branched,<br />

incurved hyphae (Fig. 25.23b). Yet another way<br />

of entrapping air within the propagule is<br />

shown by Clathrosphaerina zalewskii which forms<br />

hollow, spherical propagules with a lattice wall,<br />

resembling practice golf balls. These clathrate<br />

structures are formed by the repeated dicho<strong>to</strong>my<br />

of the arms of the developing conidium, which<br />

then curve inwards and join firmly where<br />

the tips of the arms <strong>to</strong>uch (Figs. 25.22b, 25.24).<br />

This fungus has a minute inoperculate discomycete<br />

teleomorph, a species of Hyaloscypha<br />

whose apothecia develop in air on twigs or<br />

pieces of wood which have previously been<br />

submerged.<br />

Several bubble-trap propagules are also<br />

known among Basidiomycota. An example<br />

is Aegerita candida (teleomorph Bulbillomyces<br />

farinosus), which forms its conidia (bulbils)<br />

and basidia on the surface of wet, previously<br />

submerged wood. The propagule is made up<br />

of tightly clustered aggregates of inflated,<br />

dikaryotic clamped cells between which air is<br />

entrapped (see Figs. 18.8 and 25.22c). The<br />

propagules of Aegeritina <strong>to</strong>rtuosa (teleomorph<br />

Subulicystidium longisporum) which also grows on<br />

wet wood resemble those of A. candida but are<br />

not clamped.


698 ANAMORPHIC FUNGI<br />

Fig 25.21 Some aero-aquatic helicosporous fungi. (a) Helicoon<br />

richonis.(b)Helicodendron triglitziense.(c)Helicodendron<br />

conglomeratum.The central spore is drawn in optical section <strong>to</strong><br />

show the trapped air bubble.<br />

25.3.2 Ecophysiological studies<br />

Although aero-aquatic fungi are taxonomically<br />

diverse, they share several common physiological<br />

features which help in understanding their<br />

ecology (for references see Webster & Descals,<br />

1981; Dix & Webster, 1995; Voglmayr, 2000).<br />

Simple techniques have aided studies of their<br />

ecology. Quantitative studies on colonization<br />

and survival have been made using small discs<br />

of leaves of beech, Fagus sylvatica. Sterile or<br />

artificially inoculated discs can be submerged<br />

among the accumulated leaf detritus in a pond<br />

and recovered at intervals in order <strong>to</strong> moni<strong>to</strong>r<br />

the development of conidia for quantifying<br />

colonization.<br />

The bubble-trap propagules of aero-aquatic<br />

fungi are hydrophobic and float ungerminated<br />

at the water surface of stagnant ponds. Autumn-


AERO-AQUATIC FUNGI<br />

699<br />

Fig 25.22 Propagules of aero-aquatic fungi which have<br />

developed at the surface of moist leaves incubated in air.<br />

(a) Helicodendron giganteum. Note that secondary conidia<br />

can develop as branches from the first-formed conidia.<br />

(b) Clathrosphaerinazalewskii.(c)Aegerita candida. All images<br />

<strong>to</strong> same scale.Reprinted from Dix and Webster (1995); original<br />

micrographs kindly provided by P. J. Fisher.<br />

shed leaves which fall on<strong>to</strong> the water are rapidly<br />

colonized (Premdas & Kendrick, 1991). The leaves<br />

eventually sink <strong>to</strong> the bot<strong>to</strong>m of the pond and<br />

growth of the fungi continues so long as<br />

dissolved oxygen is available in the water.<br />

Underwater colonization may also take place<br />

by leaf-<strong>to</strong>-leaf contact, shown by the fact that<br />

sterilized beech leaf discs submerged among<br />

detritus develop conidia of aero-aquatic fungi<br />

when later incubated under suitable conditions<br />

out of water. If there is a substantial accumulation<br />

of fallen leaves, the metabolic activity of<br />

decomposer organisms will lead <strong>to</strong> anaerobic<br />

conditions and the evolution of hydrogen<br />

sulphide (H 2 S) which is <strong>to</strong>xic at low concentrations<br />

<strong>to</strong> eukaryotic organisms. Anaerobic conditions<br />

are evidenced by the sulphurous smell of<br />

disturbed mud and by the black colour of the silt<br />

caused by the accumulation of metallic<br />

sulphides, especially iron sulphide. Although<br />

aero-aquatic hyphomycetes grow best at atmospheric<br />

oxygen levels, their growth in anoxic<br />

conditions is still superior <strong>to</strong> that of other fungi<br />

(Fisher & Webster, 1979). Under strictly anaerobic<br />

conditions, five species of Helicodendron<br />

showed almost 100% survival for 6 months, and<br />

substantial survival even after 12 months (Field &<br />

Webster, 1983). Survival in most cases appears <strong>to</strong><br />

be by thick-walled hyphae because chlamydospores<br />

and sclerotia are rarely found. Similar<br />

comparative studies of the survival of aeroaquatic<br />

fungi and aquatic hyphomycetes under<br />

anaerobic conditions in the presence of low<br />

concentrations of H 2 S showed better survival of<br />

aero-aquatic fungi than aquatic hyphomycetes<br />

(Field & Webster, 1985).<br />

Following a prolonged period of submersion<br />

under the anaerobic or near-anaerobic conditions<br />

of the bot<strong>to</strong>m silt of a pond, it takes several<br />

days for sporulation of aero-aquatic fungi <strong>to</strong><br />

commence in air. Incubation in well-aerated<br />

water before exposure <strong>to</strong> air improves the<br />

recovery, suggesting that a period of aerobic<br />

growth is a stimulus <strong>to</strong> sporulation. For most<br />

species studied, exposure <strong>to</strong> light also enhances<br />

sporulation (Fisher & Webster, 1978).<br />

Aero-aquatic fungi (Helicodendron spp.) grown<br />

on a homogenized beech leaf mash can survive<br />

on the soil surface for several months if airdried<br />

(Fisher, 1978), suggesting a capacity for<br />

vegetative existence out of water, a conclusion<br />

confirmed by reports of some species from soil<br />

(Abdullah & Webster, 1980). Dispersal of colonized<br />

wind-blown leaves from one body of water<br />

<strong>to</strong> another is clearly a possibility. Other possible


700 ANAMORPHIC FUNGI<br />

Fig 25.23 Air-trapping<br />

multicellular propagules of<br />

two aero-aquatic fungi.<br />

(a) Beverwijkella pulmonaria.<br />

(b) Spirosphaera sp.<br />

Fig 25.24<br />

Clathrosphaerina zalewskii.Conidial development.


AERO-AQUATIC FUNGI<br />

701<br />

means of dispersal are the carriage of wind-borne<br />

ascospores and basidiospores of the teleomorphs,<br />

passive dispersal of conidia by waterfowl, and<br />

conidium dispersal at the water surface during<br />

flooding.<br />

Aero-aquatic fungi show some host specificity.<br />

Leaves and twigs of broad-leaved trees are<br />

colonized by a wide range of species, but leaves<br />

of coniferous needles such as Pinus have a more<br />

restricted mycota including Hd. fractum and<br />

Hd. hyalinum. Monocotyledonous hosts such as<br />

grasses, rushes and sedges support few species,<br />

but exceptions are Hd. praetermissum, Spirosphaera<br />

carici-graminis and S. minuta. Fruiting of Aegerita<br />

and Aegeritina is supported on woody substrates<br />

rather than leaves, whereas Pseudaegerita spp.<br />

fruit both on rotten wood and on leaves. There<br />

are also differences in frequency in relation <strong>to</strong><br />

water chemistry (Voglmayr, 2000). For example<br />

B. pulmonaria, Hd. conglomeratum, Hd. tubulosum<br />

and Helicoon fuscosporum are found in eutrophic<br />

waters, Spirosphaera minuta is characteristic<br />

of oligotrophic conditions, and Candelabrum<br />

desmidiaceum is mostly found in dystrophic<br />

conditions. Other species have a wider <strong>to</strong>lerance.<br />

Helicodendron triglitziense and S. floriformis are<br />

two very common species found in dystrophic<br />

<strong>to</strong> eutrophic waters.<br />

Like so many fungi described in this book,<br />

aero-aquatic species have adopted, by convergent<br />

evolution, a suite of flexible characters which<br />

has enabled them <strong>to</strong> explore a habitat of<br />

extremes, in this case the amphibious environment<br />

involving changes between anaerobic and<br />

aerobic conditions, and between aquatic and<br />

terrestrial lifestyles.


References<br />

Abbott, S. P. & Currah, R. S. (1997). The Helvellaceae:<br />

systematic revision and occurrence in northwestern<br />

North America. Mycotaxon, 62, 1 125.<br />

Abbott, S. P. & Sigler, L. (2001). Heterothallism in the<br />

Microascaceae demonstrated by three species in the<br />

Scopulariopsis brevicaulis series. Mycologia, 93,<br />

1211 1220.<br />

Abdullah, S. K. (1980). British aero-aquatic fungi Ph.D.<br />

thesis, University of Exeter, UK.<br />

Abdullah, S. K. & Webster, J. (1980). Occurrence of<br />

aero-aquatic fungi in soil. Transactions of the British<br />

Mycological Society, 75, 511 514.<br />

Adam, G. C. (1988). Thanatephorus cucumeris (Rhizoc<strong>to</strong>nia<br />

solani), a species of wide host range. Advances in Plant<br />

Pathology, 6, 535 552.<br />

Adams, M. J. (1991). Transmission of plant viruses by<br />

fungi. Annals of Applied Biology, 118, 479 492.<br />

Adams, T. H. & Yu, J.-H. (1998). Coordinate control of<br />

secondary metabolite production and asexual<br />

sporulation in Aspergillus nidulans. Current Opinion in<br />

Microbiology, 1, 674 677.<br />

Adams, T. H., Wieser, J. K. & Yu, J.-H. (1998). Asexual<br />

sporulation in Aspergillus nidulans. Microbiology and<br />

Molecular Biology Reviews, 62, 35 54.<br />

Adams, T. J. H., Todd, N. K. & Rayner, A. D. M. (1981).<br />

Antagonism between dikaryons of Pip<strong>to</strong>porus<br />

betulinus. Transactions of the British Mycological Society,<br />

76, 510 513.<br />

Adendorff, R. & Rijkenberg, F. H. J. (2000). Scanning<br />

electron microscopy of direct host leaf penetration<br />

by urediospore-derived infection structures of<br />

Phakopsora apoda. Mycological Research, 104, 317 324.<br />

Adler, E. (1977). Lignin chemistry. Past, present<br />

and future. Wood Science and Technology, 11,<br />

169 218.<br />

Agerer, R. (2001). Exploration types of ec<strong>to</strong>mycorrhizae.<br />

A proposal <strong>to</strong> classify ec<strong>to</strong>mycorrhizal mycelial<br />

systems according <strong>to</strong> their patterns of differentiation<br />

and putative ecological importance. Mycorrhiza,<br />

11, 107 114.<br />

Agerer, R. (2002). Rhizomorph structures confirm the<br />

relationship between Lycoperdales and Agaricaceae<br />

(Hymenomycetes, Basidiomycota). Nova Hedwigia, 75,<br />

367 385.<br />

Agrios, G. N. (2005). Plant Pathology, 5th edn.<br />

Amsterdam: Elsevier.<br />

Ahmadjian, V. (1993). The Lichen Symbiosis. New York:<br />

John Wiley.<br />

Ahmadjian, V. & Jacobs, J. B. (1981). Relationship<br />

between fungus and alga in the lichen Cladonia<br />

cristatella Tuck. Nature, 289, 169 172.<br />

Åhman, J., Ek, B., Rask, L. & Tunlid, A. (1996). Sequence<br />

analysis and regulation of a cuticle-degrading serine<br />

protease from the nema<strong>to</strong>phagous fungus<br />

Arthrobotrys oligospora. Microbiology, 142, 1605 1616.<br />

Åhman, J., Johansson, T., Olsson, M., et al. (2002).<br />

Improving the pathogenicity of a nema<strong>to</strong>de-trapping<br />

fungus by genetic engineering of a subtilisin with<br />

nema<strong>to</strong><strong>to</strong>xic activity. Applied and Environmental<br />

Microbiology, 68, 3408 3415.<br />

Ahmed, S. I. & Cain, R. F. (1972). Revision of the genera<br />

Sporormia and Sporormiella. Canadian Journal of Botany,<br />

50, 419 477.<br />

Aidoo, K. E., Smith, J. E. & Wood, B. J. B. (1994).<br />

Industrial aspects of soy sauce fermentations using<br />

Aspergillus. InThe Genus Aspergillus. From Taxonomy and<br />

Genetics <strong>to</strong> Industrial Application, ed. K. A. Powell,<br />

A. Renwick & J. F. Peberdy. New York: Plenum Press,<br />

pp. 155 169.<br />

Ainsworth, A. M. (1987). Occurrence and interactions of<br />

outcrossing and non-outcrossing populations of<br />

Stereum, Phanerochaete and Coniophora. In Evolutionary<br />

Biology of <strong>Fungi</strong>, ed. A. D. M. Rayner, C. M. Brasier &<br />

D. Moore. Cambridge: Cambridge University Press,<br />

pp. 285 299.<br />

Ainsworth, A. M. & Rayner, A. D. M. (1989). Hyphal and<br />

mycelial responses associated with genetic exchange<br />

within and between species of the basidiomycete<br />

genus Stereum. Journal of General Microbiology, 135,<br />

1643 1659.<br />

Ainsworth, A. M., Rayner, A. D. M., Broxholme, S. J. &<br />

Beeching, J. R. (1990). Occurrence of unilateral<br />

genetic transfer and genomic replacement between<br />

strains of Stereum hirsutum from non-outcrossing and<br />

outcrossing populations. New Phy<strong>to</strong>logist, 115,<br />

119 128.<br />

Ainsworth, G. C. (1973). <strong>Introduction</strong> and keys <strong>to</strong><br />

higher taxa. In The <strong>Fungi</strong>. An Advanced Treatise IVB: A<br />

Taxonomic Review with Keys, ed. G. C. Ainsworth,<br />

F. K. Sparrow & A. S. Sussman. New York: Academic<br />

Press, pp. 1 7.<br />

Ainsworth, G. C. (1976). <strong>Introduction</strong> <strong>to</strong> the His<strong>to</strong>ry of<br />

Mycology. Cambridge: Cambridge University Press.<br />

Ainsworth, G. C. (1981). <strong>Introduction</strong> <strong>to</strong> the His<strong>to</strong>ry of<br />

Plant Pathology. Cambridge: Cambridge University<br />

Press.


REFERENCES<br />

703<br />

Aist, J. R. & Israel, H. W. (1977). Timing and significance<br />

of papilla formation during host penetration by<br />

Olpidium brassicae. Phy<strong>to</strong>pathology, 67, 187 194.<br />

Aist, J. R. & Williams, P. H. (1971). The cy<strong>to</strong>logy and<br />

kinetics of cabbage root hair penetration by<br />

Plasmodiophora brassicae. Canadian Journal of Botany,<br />

49, 2023 2034.<br />

Akashi, T., Kanbe, T. & Tanaka, K. (1997). Localized<br />

accumulation of cell wall mannoproteins and<br />

endomembranes induced by brefeldin A in hyphal<br />

cells of the dimorphic yeast Candida albicans.<br />

Pro<strong>to</strong>plasma, 197, 45 56.<br />

Akins, R. A. (2005). An update on antifungal targets and<br />

mechanisms of resistance in Candida. Medical<br />

Mycology, 43, 285 318.<br />

Alasoadura, S. O. (1963). Fruiting in Sphaerobolus with<br />

special reference <strong>to</strong> light. Annals of Botany N. S., 27,<br />

123 145.<br />

Alberts, B., Johnson, A., Lewis, J., et al. (2002).<br />

Molecular Biology of the Cell, 4th edn. New York:<br />

Garland Science.<br />

Alcorn, J. L. (1981). Ascus structure and function in<br />

Cochliobolus species. Mycotaxon, 13, 349 360.<br />

Alcorn, J. L. (1988). The taxonomy of ‘Helminthosporium’<br />

species. Annual Review of Plant Pathology, 26, 37 56.<br />

Alderman, D. J. & Polglase, J. L. (1986). Aphanomyces<br />

astaci: isolation and culture. Journal of Fish Diseases, 9,<br />

367 379.<br />

Alderman, D. J., Holdich, D. & Reeve, I. (1990). Signal<br />

crayfish as vec<strong>to</strong>rs in crayfish plague in Britain.<br />

Aquaculture, 86, 3 6.<br />

Alderman, S. C. (2003). Diversity and speciation in<br />

Claviceps. In Clavicipitalean <strong>Fungi</strong>, ed. J. F. White,<br />

C. W. Bacon, N. L. Hywel-Jones & J. W. Spatafora.<br />

New York: Marcel Dekker Inc., pp. 190 245.<br />

Alexander, C. & Hadley, G. (1985). Carbon movement<br />

between host and mycorrhizal endophyte during<br />

development of the orchid Goodyera repens Br. New<br />

Phy<strong>to</strong>logist, 101, 657 665.<br />

Alexander, H. M. & An<strong>to</strong>novics, J. (1988). Disease spread<br />

and population dynamics of anther-smut infection<br />

of Silene alba caused by the fungus Ustilago violacea.<br />

Journal of Ecology, 76, 91 104.<br />

Alexopoulos, C. J., Mims, C. W. & Blackwell, M. (1996).<br />

Introduc<strong>to</strong>ry Mycology, 4th edn. New York: Wiley.<br />

Allen, E. A., Hoch, H. C., Stavely, J. R. & Steadman, J. R.<br />

(1991). Uniformity among races of Uromyces appendiculatus<br />

in response <strong>to</strong> <strong>to</strong>pographic signaling for<br />

appressorium formation. Phy<strong>to</strong>pathology, 81,<br />

883 887.<br />

Allen, M. F. (1991). The Ecology of Mycorrhizae. Cambridge:<br />

Cambridge University Press.<br />

Allen, M. F., Klironomos, J. N. & Harney, S. (1997). The<br />

epidemiology of mycorrhizal fungal infection<br />

during succession. In The Mycota VB: Plant<br />

Relationships, ed. G. C. Carroll & P. Tudzynski. Berlin:<br />

Springer-Verlag, pp. 169 83.<br />

Almaraz, T., Roux, C., Maumont, S. & Durrieu, G.<br />

(2002). Phylogenetic relationships among smut fungi<br />

parasitizing dicotyledons based on ITS sequence<br />

analysis. Mycological Research, 106, 541 548.<br />

Amagai, A. (1992). Induction of heterothallic and<br />

homothallic zygotes in Dictyostelium discoideum by<br />

ethylene. Development Growth and Differentiation, 34,<br />

293 299.<br />

Amano, K. (1986). Host Range and Geographical<br />

Distribution of the Powdery Mildew <strong>Fungi</strong>. Tokyo: Japan<br />

Scientific Societies Press.<br />

Amon, J. P. & French, K. H. (2004). Pho<strong>to</strong>responses of<br />

the marine protist Ulkenia sp. zoospores <strong>to</strong> ambient,<br />

artificial and bioluminescent light. Mycologia, 96,<br />

463 469.<br />

Amon, J. P. & Perkins, F. O. (1968). Structure of<br />

Labyrinthula sp. zoospores. Journal of Pro<strong>to</strong>zoology, 15,<br />

543 546.<br />

Anagnostakis, S. L. (1987). Chestnut blight: the classical<br />

problem of an introduced pathogen. Mycologia, 79,<br />

23 37.<br />

Anagnostakis, S. L. & Day, P. R. (1979). Hypovirulence<br />

conversion in Endothia parasitica. Phy<strong>to</strong>pathology, 69,<br />

1226 1229.<br />

Andersen, R. A., Barr, D. J. S., Lynn, D. H., et al. (1991).<br />

Terminology and nomenclature of the cy<strong>to</strong>skeletal<br />

elements associated with the flagellar/ciliary<br />

apparatus in protists. Pro<strong>to</strong>plasma, 164, 1 8.<br />

Andersen, T. F. & Stalpers, J. A. (1994). A check-list of<br />

Rhizoc<strong>to</strong>nia epithets. Mycotaxon, 51, 437 457.<br />

Anderson, D. L., Giacon, H. & Gibson, N. L. (1997).<br />

Detection and thermal destruction of the<br />

chalkbrood fungus (Ascosphaera apis) in honey.<br />

Journal of Apicultural Research, 36, 163 168.<br />

Anderson, D. L., Gibbs, A. J. & Gibson, N. L. (1998).<br />

Identification and phylogeny of spore cyst fungi<br />

(Ascosphaera spp.) using ribosomal DNA sequences.<br />

Mycological Research, 102, 541 547.<br />

Anderson, I. C., Chambers, S. M. & Cairney, J. W. G.<br />

(1998). Use of molecular methods <strong>to</strong> estimate the<br />

size and distribution of mycelial individuals of the<br />

ec<strong>to</strong>mycorrhizal basidiomycete Pisolithus tinc<strong>to</strong>rius.<br />

Mycological Research, 102, 295 300.<br />

Anderson, I. C., Chambers, S. M. & Cairney, J. W. G.<br />

(2001). Distribution and persistence of Australian<br />

Pisolithus species genets at native sclerophyll forest<br />

field sites. Mycological Research, 105, 971 976.


704 REFERENCES<br />

Anderson, J. B. & Ullrich, R. C. (1982). Diploids of<br />

Armillaria mellea: synthesis, stability, and mating<br />

behavior. Canadian Journal of Botany, 60, 432 439.<br />

Anderson, J. B., Ullrich, R. C., Roth, L. F. & Filip, G. M.<br />

(1979). Genetic identification of clones of Armillaria<br />

mellea in coniferous forests in Washing<strong>to</strong>n.<br />

Phy<strong>to</strong>pathology, 69, 1109 1111.<br />

Ando, K. & Tubaki, K. (1984a). Some undescribed<br />

hyphomycetes in the raindrops from intact leaf<br />

surfaces. Transactions of the Mycological Society of Japan,<br />

25, 21 37.<br />

Ando, K. & Tubaki, K. (1984b). Some undescribed<br />

hyphomycetes in rainwater draining from intact<br />

trees. Transactions of the Mycological Society of Japan, 25,<br />

39 47.<br />

Andrianopoulos, A. (2002). Control of morphogenesis<br />

in the human fungal pathogen Penicillium marneffei.<br />

International Journal of Medical Microbiology, 292,<br />

331 347.<br />

Andrivon, D. (1995). Biology, ecology, and epidemiology<br />

of the pota<strong>to</strong> late blight pathogen Phy<strong>to</strong>phthora<br />

infestans in soil. Phy<strong>to</strong>pathology, 85, 1053 1056.<br />

Anikster, Y. (1983). Binucleate basidiospores a<br />

general rule in rust fungi. Transactions of the British<br />

Mycological Society, 81, 624 626.<br />

Anikster, Y. (1984). The formae speciales. InThe Cereal<br />

Rusts 1: Origins, Specificity, Structure, and Physiology, ed.<br />

W. R. Bushnell & A. P. Roelfs. Orlando: Academic<br />

Press, pp. 115 130.<br />

Anikster, Y. & Wahl, I. (1979). Coevolution of the rust<br />

fungi on Gramineae and Liliaceae and their hosts.<br />

Annual Review of Phy<strong>to</strong>pathology, 17, 367 403.<br />

Anikster, Y. & Wahl, I. (1985). Basidiospore formation<br />

and self-fertility in Puccinia mesnieriana. Transactions<br />

of the British Mycological Society, 84, 164 167.<br />

Anikster, Y., Bushnell, W. R., Eilam, T., Manisterski, J. &<br />

Roelfs, A. P. (1997). Puccinia recondita causing leaf rust<br />

on cultivated wheats, wild wheats, and rye. Canadian<br />

Journal of Botany, 75, 2082 2096.<br />

Anikster, Y., Eilam, T., Manisterski, J. & Leonard, K. J.<br />

(2003). Self-fertility and other distinguishing<br />

characteristics of a new morphotype of Puccinia<br />

coronata pathogenic on smooth brome grass.<br />

Mycologia, 95, 87 97.<br />

Anke, H., Stadler, M., Mayer, A. & Sterner, O. (1995).<br />

Secondary metabolites with nematicidal and antimicrobial<br />

activity from nema<strong>to</strong>phagous fungi and<br />

Ascomycetes. Canadian Journal of Botany, 73,<br />

S932 S939.<br />

Anke, T. (1997). Strobilurins. In Fungal Biotechnology,<br />

ed. T. Anke. Weinheim, Germany: Chapman & Hall,<br />

pp. 206 212.<br />

Anke, T., Oberwinkler, F., Steglich, W. & Schramm, G.<br />

(1977). The strobilurins new antifungal<br />

antibiotics from the basidiomycete Strobilurus<br />

tenacellus (Pers. Ex Fr.) Sing. Journal of Antibiotics,<br />

30, 806 810.<br />

Ann, P.-J., Chang, T.-T. & Ko, W.-H. (2002). Phellinus noxius<br />

brown root rot of fruit and ornamental trees in<br />

Taiwan. Plant Disease, 86, 820 826.<br />

Ansell, P. J. & Young, T. W. K. (1982). Light and electronmicroscopic<br />

study of Mortierella species. Nova<br />

Hedwigia, 36, 309 339.<br />

Ansell, P. J. & Young, T. W. K. (1983). Light and electron<br />

microscopy of Mortierella indohi zygospores. Mycologia,<br />

75, 64 69.<br />

Ansell, P. J. & Young, T. W. K. (1988). The zygospore of<br />

Mortierella indohi. Transactions of the British Mycological<br />

Society, 91, 221 226.<br />

Antelo, L., Cosio, E. G., Hertkorn, N. & Ebel, J. (1998).<br />

Partial purification of a GTP-insensitive (1!3)-bglucan<br />

synthase from Phy<strong>to</strong>phthora sojae. FEBS Letters,<br />

433, 191 195.<br />

An<strong>to</strong>novics, J., Hood, M. & Partain, J. (2002). The<br />

ecology and genetics of a host shift:<br />

Microbotryum as a model system. American<br />

Naturalist, 160, S40 S53.<br />

Arabi, M. I. E. & Jawhar, M. (2002). The ability of barley<br />

powdery mildew <strong>to</strong> grow in vitro. Journal of<br />

Phy<strong>to</strong>pathology, 150, 305 307.<br />

Archibald, J. M. & Keeling, P. J. (2004). Actin and<br />

ubiquitin protein sequences support a cercozoan/<br />

foraminiferan ancestry for the plasmodiophorid<br />

plant pathogens. Journal of Eukaryotic Microbiology, 51,<br />

113 118.<br />

Arditti, J. H. (1992). Fundamentals of Orchid Biology.<br />

New York: John Wiley & Sons.<br />

Arellano, M., Cartagena-Lirola, H., Hajibagheri, M. A. N.,<br />

Durán, A. & Valdivieso, M. H. (2000). Proper ascospore<br />

maturation requires the chs þ chitin synthase gene in<br />

Schizosaccharomyces pombe. Molecular Microbiology, 35,<br />

79 89.<br />

Arenal, F., Platas, G., Monte, E. & Peláez, F. (2000). ITS<br />

sequencing support for Epicoccum nigrum and Phoma<br />

epicoccina being the same biological species.<br />

Mycological Research, 104, 301 303.<br />

Aristizabal, B. H., Clemons, K. V., Stevens, D. A. &<br />

Restrepo, A. (1998). Morphological transition of<br />

Paracoccidioides brasiliensis conidia <strong>to</strong> yeast cells: in<br />

vivo inhibition in females. Infection and Immunity, 66,<br />

5587 5591.<br />

Arnold, C. A. (1928). The development of the<br />

perithecium and spermagonium of Sporormia<br />

leporina. American Journal of Botany, 15, 241 245.


REFERENCES<br />

705<br />

Arnold, D. L., Blakesley, D. & Clarkson, J. M. (1996).<br />

Evidence for the growth of Plasmodiophora brassicae in<br />

vitro. Mycological Research, 100, 535 540.<br />

Arroyo, M., de la Mata, I., Acebal, C. & Castillón, M. P.<br />

(2003). Biotechnological applications of penicillin<br />

acylases: state-of-the-art. Applied Microbiology and<br />

Biotechnology, 60, 507 514.<br />

Arsuffi, T. L. & Suberkropp, K. (1985). Selective feeding<br />

by stream caddisfly (Trichoptera) detritivores on<br />

leaves with fungal-colonized patches. Oikos, 45,<br />

50 58.<br />

Aschner, M. & Kohn, S. (1958). The biology of<br />

Harposporium anguillulae. Journal of General<br />

Microbiology, 19, 182 189.<br />

Ashbee, H. R. & Evans, E. G. V. (2002). Immunology of<br />

diseases associated with Malassezia species. Clinical<br />

Microbiology Reviews, 15, 21 57.<br />

Ashford, A. E. & Allaway, W. G. (2002). The role of the<br />

motile tubular vacuole system in mycorrhizal fungi.<br />

Plant and Soil, 244, 177 187.<br />

Ashford, A. E., Cole, L. & Hyde, G. J. (2001). Motile<br />

tubular vacuole systems. In The Mycota VIII: Biology of<br />

the Fungal Cell, ed. R. J. Howard & N. A. R. Gow. Berlin:<br />

Springer-Verlag, pp. 243 265.<br />

Asiegbu, F. O., Adomas, A. & Stenlid, J. (2005). Conifer<br />

root and butt rot caused by Heterobasidion annosum<br />

(Fr.) Bref. s.l. Molecular Plant Pathology, 6, 395 409.<br />

Asiegbu, F. O., Johansson, M., Woodward, S. &<br />

Hütterman, A. (1998). Biochemistry of the<br />

host parasite interaction. In Heterobasidion annosum.<br />

Biology, Ecology, Impact and Control, ed. S. Woodward, J.<br />

Stenlid, R. Karjalainen & A. Hütterman. Wallingford,<br />

UK: CAB International, pp. 167 193.<br />

Au, D. W. T., Jones, E. B. G., Moss, S. T. & Hodgkiss, I. J.<br />

(1996). The role of mucilage in the attachment of<br />

conidia, germ tubes and appressoria in the saprobic<br />

aquatic hyphomycetes Lemonniera aquatica and<br />

Mycocentrospora filiformis. Canadian Journal of Botany,<br />

74, 1789 1800.<br />

Aust, H.-J. (1981). Über den Verlauf von<br />

Mehltauepidemien innerhalb des Agrar-Ökosystems<br />

Gerstenfeld. Acta Phy<strong>to</strong>medica, 7, 1 76.<br />

Austin, D. J., Bu’lock, J. D. & Gooday, G. W. (1969).<br />

Trisporic acids: sexual hormones for Mucor mucedo<br />

and Blakeslea trispora. Nature, 223, 1178 1179.<br />

Austwick, P. C. K. (1976). Environmental aspects of<br />

Mortierella wolfii infection in cattle. New Zealand<br />

Journal of Agricultural Research, 19, 25 33.<br />

Avis, T. J. & Bélanger, R. R. (2002). Mechanisms and<br />

means of detection of biocontrol activity of<br />

Pseudozyma yeasts against plant-pathogenic fungi.<br />

FEMS Yeast Research, 2, 5 8.<br />

Ayers, W. A. & Lumsden, R. D. (1977). Mycoparasitism of<br />

oospores of Pythium and Aphanomyces species by<br />

Hyphochytrium catenoides. Canadian Journal of<br />

Microbiology, 23, 38 44.<br />

Aylmore, R. C., Wakley, G. E. & Todd, N. K. (1984). Septal<br />

sealing in the basidiomycete Coriolus versicolor. Journal<br />

of General Microbiology, 130, 2975 2982.<br />

Baba, M. & Osumi, M. (1987). Transmission and<br />

scanning electron microscopic examination of<br />

intracellular organelles in freeze-substituted<br />

Kloeckera and Saccharomyces cerevisiae yeast cells.<br />

Journal of Electron Microscopy Technique, 5, 249 261.<br />

Backhouse, D. & Willetts, H. J. (1985). His<strong>to</strong>chemical<br />

changes during conidiogenic germination of sclerotia<br />

of Botrytis cinerea. Canadian Journal of Microbiology,<br />

31, 282 286.<br />

Bacon, C. W. & Hin<strong>to</strong>n, D. M. (1988). Ascospore iterative<br />

germination in Epichloe typhina. Transactions of the<br />

British Mycological Society, 90, 563 569.<br />

Bacon, C. W. & Hin<strong>to</strong>n, D. M. (1991). Microcyclic<br />

conidiation cycles in Epichloe typhina. Mycologia, 83,<br />

743 751.<br />

Bacon, C. W., Porter, J. K., Robbins, J. D. & Luttrell, E. S.<br />

(1977). Epichloe typhina from <strong>to</strong>xic tall fescue grasses.<br />

Applied and Environmental Microbiology, 34, 576 581.<br />

Bacon, C. W., Richardson, M. D. & White, J. F. (1997).<br />

Modification and uses of endophyte-enhanced turfgrasses.<br />

Crop Science, 37, 1415 1425.<br />

Bagchee, K. (1954). Merulius lacrymans (Wulf.) Fr. in<br />

India. Sydowia, 8, 80 85.<br />

Bähler, J., Steever, A. B., Wheatley, S., et al. (1998). Role<br />

of Polo kinase and Mid1p in determining the site of<br />

cell division in fission yeast. Journal of Cell Biology,<br />

143, 1603 1616.<br />

Bai, F.-Y., Zhao, J.-H., Takashima, M., et al. (2002).<br />

Reclassification of the Sporobolomyces roseus and the<br />

Sporidiobolus pararoseus complexes, with the<br />

description of Sporobolomyces phaffii sp. nov.<br />

International Journal of Systematic and Evolutionary<br />

Microbiology, 52, 2309 2314.<br />

Bailey, F. & Gray, N. F. (1989). The comparison of<br />

isolation techniques for nema<strong>to</strong>phagous<br />

fungi from soil. Annals of Applied Biology, 114,<br />

125 132.<br />

Bailey, J. A. & Jeger, M. J., eds. (1992). Colle<strong>to</strong>trichum:<br />

Biology, Pathology and Control. Wallingford, UK: CAB<br />

International.<br />

Bailey, J. A., O’Connell, R. J., Pring, R. J. & Nash, C. (1992).<br />

Infection strategies of Colle<strong>to</strong>trichum species. In<br />

Colle<strong>to</strong>trichum: Biology, Pathology and Control, ed.<br />

J. A. Bailey & M. J. Jeger. Wallingford, UK: CAB<br />

International, pp. 88 120.


706 REFERENCES<br />

Baka, Z. A., Larous, L. & Lösel, D. M. (1995). Distribution<br />

of ATPase activity at the host pathogen interfaces of<br />

rust infections. Physiological and Molecular Plant<br />

Pathology, 47, 67 82.<br />

Bakkeren, G., Kronstad, J. W. & Lévesque, C. A. (2000).<br />

Comparison of AFLP fingerprints and ITS sequences<br />

as phylogenetic markers in Ustilaginomycetes.<br />

Mycologia, 92, 510 521.<br />

Balazy, S. (1984). On rhizoids of En<strong>to</strong>mophthora muscae<br />

(Cohn) Fresenius (En<strong>to</strong>mophthorales:<br />

En<strong>to</strong>mophthoraceae). Mycotaxon, 19, 397 407.<br />

Baldauf, S. L. (1999). A search for the origins of<br />

animals and fungi: comparing and combining<br />

molecular data. American Naturalist, 154,<br />

S178 S188.<br />

Baldauf, S. L., Roger, A. J., Wenk-Siefert, I. & Doolittle,<br />

W. F. (2000). A kingdom-level phylogeny of eukaryotes<br />

based on combined protein data. Science, 290,<br />

972 977.<br />

Bandoni, R. J. (1963). Conjugation in Tremella mesenterica.<br />

Canadian Journal of Botany, 41, 467 474.<br />

Bandoni, R. J. (1965). Secondary control of conjugation<br />

in Tremella mesenterica. Canadian Journal of Botany, 43,<br />

627 630.<br />

Bandoni, R. J. (1972). Terrestrial occurrence of some<br />

aquatic hyphomycetes. Canadian Journal of Botany, 50,<br />

2283 2288.<br />

Bandoni, R. J. (1974). Mycological observations on the<br />

aqueous films covering decaying leaves and other<br />

litter. Transactions of the Mycological Society of Japan, 15,<br />

309 315.<br />

Bandoni, R. J. (1984). The Tremellales and<br />

Auriculariales: an alternative classification.<br />

Transactions of the Mycological Society of Japan, 25,<br />

489 530.<br />

Bandoni, R. J. (1987). Taxonomic overview of the<br />

Tremellales. Studies in Mycology, 30, 87 110.<br />

Bandoni, R. J. & Boekhout, T. (1998). Tremelloid genera<br />

with yeast phases. In The Yeasts. A Taxonomic Study, 4th<br />

edn, ed. C. P. Kurtzman & J. W. Fell. Amsterdam:<br />

Elsevier, pp. 705 717.<br />

Bandoni, R. J., Lobo, K. J. & Brezden, S. A. (1971).<br />

Conjugation and chlamydospores in<br />

Sporobolomyces odorus. Canadian Journal of Botany,<br />

49, 683 686.<br />

Banuett, F. (1995). Genetics of Ustilago maydis, a fungal<br />

pathogen that induces tumours in maize. Annual<br />

Review of Genetics, 29, 179 208.<br />

Banuett, F. & Herskowitz, I. (1996). Discrete developmental<br />

stages during teliospore formation in the<br />

corn smut fungus, Ustilago maydis. Development, 122,<br />

2965 2976.<br />

Bao, J. R. & Lasarovits, G. (2001). Differential colonization<br />

of <strong>to</strong>ma<strong>to</strong> roots by nonpathogenic and<br />

pathogenic Fusarium oxysporum strains may influence<br />

fusarium wilt control. Phy<strong>to</strong>pathology, 91, 449 456.<br />

Bardwell, L. (2004). A walk-through of the yeast mating<br />

pheromone response pathway. Peptides, 25,<br />

1465 1476.<br />

Barksdale, A. W. (1960). Inter-thallic sexual reactions in<br />

Achlya, a genus of the aquatic fungi. American Journal<br />

of Botany, 47, 14 23.<br />

Barksdale, A. W. (1962). Effect of nutritional deficiency<br />

on growth and sexual reproduction of Achlya<br />

ambisexualis. American Journal of Botany, 49, 633 638.<br />

Barksdale, A. W. (1963). The uptake of exogenous<br />

hormone A by certain strains of Achlya. Mycologia, 55,<br />

164 171.<br />

Barksdale, A. W. (1965). Achlya ambisexualis and a new<br />

cross-conjugating species of Achlya. Mycologia, 57,<br />

493 501.<br />

Barksdale, A. W. (1966). Segregation of sex in the<br />

progeny of a selfed heterozygote of Achlya ambisexualis.<br />

Mycologia, 58, 802 804.<br />

Barksdale, A. W. (1967). The sexual hormones of the<br />

fungus Achlya. Annals of the New York Academy of<br />

Sciences, 144, 313 319.<br />

Barksdale, A. W. (1969). Sexual hormones of Achlya and<br />

other fungi. Science, 166, 831 837.<br />

Barksdale, A. W. (1970). Nutrition and antheridiolinduced<br />

branching in Achlya ambisexualis. Mycologia,<br />

62, 411 420.<br />

Bärlocher, F., ed. (1992). The Ecology of Aquatic<br />

Hyphomycetes. Ecological Studies 94. Berlin: Springer-<br />

Verlag.<br />

Bärlocher, F. & Kendrick, W. B. (1973a). <strong>Fungi</strong> in the<br />

diet of Gammarus pseudolimnaeus (Amphipoda). Oikos,<br />

24, 295 300.<br />

Bärlocher, F. & Kendrick, W. B. (1973b). <strong>Fungi</strong> and food<br />

preference of Gammarus pseudolimnaeus. Archiv für<br />

Hydrobiologie, 72, 501 516.<br />

Bärlocher, F. & Oertli, J. J. (1978a). Colonization of<br />

conifer needles by aquatic hyphomycetes. Canadian<br />

Journal of Botany, 56, 57 62.<br />

Bärlocher, F. & Oertli, J. J. (1978b). Inhibi<strong>to</strong>rs of aquatic<br />

hyphomycetes in dead conifer needles. Mycologia, 70,<br />

964 974.<br />

Barnett, J. A., Payne, R. W. & Yarrow, D. (2000). Yeasts:<br />

Characteristics and Identification, 3rd edn. Cambridge:<br />

Cambridge University Press.<br />

Barr, D. J. S. (1980). An outline for the reclassification of<br />

the Chytridiales, and for a new order, the<br />

Spizellomycetales. Canadian Journal of Botany, 58,<br />

2380 2394.


REFERENCES<br />

707<br />

Barr, D. J. S. (1987). Rhizophlyctis rosea. In Zoosporic <strong>Fungi</strong><br />

in Teaching and Research, ed. M. S. Fuller & A. Jaworski.<br />

Athens, USA: Southeastern Publishing Corporation,<br />

pp. 30 11.<br />

Barr, D. J. S. (1988). Zoosporic plant parasites as fungal<br />

vec<strong>to</strong>rs of viruses: taxonomy and life cycles of species<br />

involved. In Viruses with Fungal Vec<strong>to</strong>rs, ed. J. I. Cooper<br />

& M. J. C. Asher. Sudbury, UK: Lavenham Press Ltd,<br />

pp. 123 137.<br />

Barr, D. J. S.(1990). Phylum Chytridiomycota. In Handbook<br />

of Pro<strong>to</strong>ctista, ed. L. Margulis, J. O. Corliss, M. Melkonian<br />

& D. J. Chapman. Bos<strong>to</strong>n: Jones & Bartlett, pp. 454 466.<br />

Barr, D. J. S. (1992). Evolution and kingdoms of organisms<br />

from the perspective of a mycologist. Mycologia,<br />

84, 1 11.<br />

Barr, D. J. S. (2001). Chytridiomycota. In The Mycota VIIA:<br />

Systematics and Evolution, ed. D. J. McLaughlin, E. G.<br />

McLaughlin & P. A. Lemke. Berlin: Springer-Verlag,<br />

pp. 93 112.<br />

Barr, D. J. S. & Allan, P. M. E. (1985). A comparison of the<br />

flagellar apparatus in Phy<strong>to</strong>phthora, Saprolegnia,<br />

Thraus<strong>to</strong>chytrium, and Rhizidiomyces. Canadian Journal<br />

of Botany, 63, 138 154.<br />

Barr, D. J. S. & Désaulniers, N. L. (1988). Precise configuration<br />

of the chytrid zoospore. Canadian Journal of<br />

Botany, 66, 869 876.<br />

Barr, D. J. S. & Hartmann, V. E. (1977). Zoospore ultrastructure<br />

of Olpidium brassicae and Rhizophlyctis rosea.<br />

Canadian Journal of Botany, 55, 1221 1235.<br />

Barr, D. J. S., Kudo, H., Jakober, K. D. & Cheng, K. J.<br />

(1989). Morphology and development of rumen<br />

fungi: Neocallimastix sp., Piromyces communis, and<br />

Orpinomyces bovis gen. nov., sp. nov. Canadian Journal<br />

of Botany, 67, 2815 2824.<br />

Barr, M. E. (1958). Life his<strong>to</strong>ry studies of Mycosphaerella<br />

tassiana and M. typhae. Mycologia, 50, 501 513.<br />

Barr, M. E. (1983). The ascomycete connection.<br />

Mycologia, 75, 1 13.<br />

Barr, M. E. (2001). Ascomycota. In The Mycota VIIA:<br />

Systematics and Evolution, ed. D. J. McLaughlin, E. G.<br />

McLaughlin & P. A. Lemke. Berlin: Springer-Verlag,<br />

pp. 161 177.<br />

Barr, M. E. & Huhndorf, S. M. (2001). Loculoascomycetes.<br />

In The Mycota VIIA: Systematics and Evolution, ed. D. J.<br />

McLaughlin, E. G. McLaughlin & P. A. Lemke. Berlin:<br />

Springer-Verlag, pp. 283 305.<br />

Barron, G. L. (1977). The Nema<strong>to</strong>de-Destroying <strong>Fungi</strong>.<br />

Guelph: Canadian Biological Publications.<br />

Barron, G. L. (1981). Preda<strong>to</strong>rs and parasites of microscopic<br />

animals. In The Biology of Conidial <strong>Fungi</strong> II, ed.<br />

G. T. Cole & B. Kendrick. New York: Academic Press,<br />

pp. 167 200.<br />

Barron, G. L. (1992). Lignolytic and cellulolytic fungi as<br />

preda<strong>to</strong>rs and parasites. In The Fungal Community. Its<br />

Organization and Role in the Ecosystem, 2nd edn, ed.<br />

G. C. Carroll & D. T. Wicklow. New York: Marcel<br />

Dekker, pp. 311 326.<br />

Barron, G. L. & Dierkes, Y. (1977). Nema<strong>to</strong>phagous<br />

fungi: Hohenbuehelia, the perfect state of<br />

Nema<strong>to</strong>c<strong>to</strong>nus. Canadian Journal of Botany, 55,<br />

3054 3062.<br />

Barron, G. L. & Szijar<strong>to</strong>, E. (1986). A new species of<br />

Olpidium parasitic in nema<strong>to</strong>de eggs. Mycologia, 78,<br />

972 975.<br />

Barron, G. L. & Thorn, R. G. (1987). Destruction of<br />

nema<strong>to</strong>des by species of Pleurotus. Canadian Journal of<br />

Botany, 65, 774 778.<br />

Barron, G. L., Cain, R. F. & Gilman, J. C. (1961). The genus<br />

Microascus. Canadian Journal of Botany, 39, 1609 1631.<br />

Barron, J. L. & Hill, E. P. (1974). Ultrastructure of<br />

zoosporogenesis in Allomyces macrogynus. Journal of<br />

General Microbiology, 80, 319 327.<br />

Barry, D., Callot, G., Janex-Favre, M. C. & Parguey-Leduc,<br />

A. (1993). Morphologie et structure des hyphes<br />

externes observées sur le péridium des Tuber à<br />

écailles: evolution au cours du développement de<br />

l’ascocarpe. Canadian Journal of Botany, 71, 609 619.<br />

Barry, D., Staun<strong>to</strong>n, S. & Callot, G. (1994). Mode of<br />

absorption of water and nutrients by ascocarps of<br />

Tuber melanosporum and Tuber aestivum. Canadian<br />

Journal of Botany, 72, 317 322.<br />

Barry, D., Jaillard, B., Staun<strong>to</strong>n, S. & Callot, G. (1995).<br />

Translocation and metabolism of phosphate following<br />

absorption by ascocarps of Tuber melanosporum<br />

and Tuber aestivum. Mycological Research, 99, 167 172.<br />

Bars<strong>to</strong>w, W. E. (1987) Catenaria anguillulae. InZoosporic<br />

<strong>Fungi</strong> in Teaching and Research, ed. M. S. Fuller & A.<br />

Jaworski. Athens, USA: Southeastern Publishing<br />

Corporation, pp. 34 35.<br />

Bars<strong>to</strong>w, W. E. & Pommerville, J. (1980). The ultrastructure<br />

of cell wall formation and of gamma<br />

particles during encystment of Allomyces macrogynus<br />

zoospores. Archiv für Mikrobiologie, 128, 179 189.<br />

Bartnicki-Garcia, S. (1968). Cell wall chemistry,<br />

morphogenesis, and taxonomy of fungi. Annual<br />

Review of Microbiology, 42, 57 69.<br />

Bartnicki-Garcia, S. (1987). The cell wall in fungal<br />

evolution. In Evolutionary Biology of the <strong>Fungi</strong>, ed.<br />

A. D. M. Rayner, C. M. Brasier & D. Moore. Cambridge:<br />

Cambridge University Press, pp. 389 403.<br />

Bartnicki-Garcia, S. (1996). The hypha: unifying thread<br />

of the fungal kingdom. In A Century of Mycology, ed.<br />

B. C. Sut<strong>to</strong>n. Cambridge: Cambridge University Press,<br />

pp. 105 133.


708 REFERENCES<br />

Bartnicki-Garcia, S. (2002). Hyphal tip growth:<br />

outstanding questions. In Molecular Biology of Fungal<br />

Development, ed. H. D. Osiewacz. New York: Marcel<br />

Dekker, pp. 29 58.<br />

Bartnicki-Garcia, S. & Wang, M. C. (1983). Biochemical<br />

aspects of morphogenesis in Phy<strong>to</strong>phthora. In<br />

Phy<strong>to</strong>phthora. Its Biology, Taxonomy, Ecology, and<br />

Pathology, ed. D. C. Erwin, S. Bartnicki-Garcia & P. H.<br />

Tsao. St. Paul: APS Press, pp. 121 137.<br />

Bartnicki-Garcia, S., Ruiz-Herrera, J. & Bracker, C. E.<br />

(1979). Chi<strong>to</strong>somes and chitin synthesis. In Fungal<br />

Walls and Hyphal Growth, ed. J. H. Burnett & A. P. J.<br />

Trinci. Cambridge: Cambridge University Press,<br />

pp. 149 168.<br />

Barve, M. P., Arie, T., Salimath, S. S., Muehlbauer, F. J. &<br />

Peever, T. L. (2003). Cloning and characterization of<br />

the mating type (MAT) locus from Ascochyta rabiei<br />

(teleomorph: Didymella rabiei) and a MAT phylogeny<br />

of legume-associated Ascochyta spp. Fungal Genetics<br />

and Biology, 39, 151 167.<br />

Bashi, E. & Fokkema, N. J. (1976). Scanning electron<br />

microscopy of Sporobolomyces roseus on wheat leaves.<br />

Transactions of the British Mycological Society, 67,<br />

500 505.<br />

Basith, M. & Madelin, M. F. (1968). Studies on the<br />

production of perithecial stromata by Cordyceps<br />

militaris in artificial culture. Canadian Journal of<br />

Botany, 46, 473 480.<br />

Basse, C. & Steinberg, G. (2004). Ustilago maydis, model<br />

system for analysis of the molecular basis of fungal<br />

pathogenicity. Molecular Plant Pathology, 5, 83 92.<br />

Basse, C. W., Lottspeich, F., Steglich, W. & Kahmann, R.<br />

(1996). Two potential indole-3-acetaldehyde<br />

dehydrogenases in the phy<strong>to</strong>pathogenic fungus<br />

Ustilago maydis. European Journal of Biochemistry,<br />

242, 648 656.<br />

Batra, L. R. (1973). Nema<strong>to</strong>sporaceae<br />

(Hemiascomycetidae): taxonomy, pathogenicity,<br />

distribution and vec<strong>to</strong>r relations. ARS Technical<br />

Bulletin, 1469, 1 71. Washing<strong>to</strong>n DC: US Department<br />

of Agriculture.<br />

Batra, R., Boekhout, T., Guého, E., et al. (2005).<br />

Malassezia Baillon, emerging clinical yeasts. FEMS<br />

Yeast Research, 5, 1101 1113.<br />

Batts, C. C. V. (1955). Observations on the infection of<br />

wheat by loose smut (Ustilago tritici (Pers.) Rostr.).<br />

Transactions of the British Mycological Society, 38,<br />

465 475.<br />

Bauer, R. (2004). Basidiomyce<strong>to</strong>us interfungal cellular<br />

interactions a synopsis. In Frontiers in Basidiomycete<br />

Mycology, ed. R. Agerer, M. Piepenbring & P. Blanz.<br />

Eching, Germany: IHW-Verlag, pp. 325 337.<br />

Bauer, R., Oberwinkler, F. & Vánky, K. (1997).<br />

Ultrastructural markers and systematics in smut<br />

fungi and allied taxa. Canadian Journal of Botany, 75,<br />

1273 1314.<br />

Bauer, R., Begerow, D., Oberwinkler, F., Piepenbring, M.<br />

& Berbee, M. L. (2001). Ustilaginomycetes. In The<br />

Mycota VIIB: Systematics and Evolution, ed. D. J.<br />

McLaughlin, E. G. McLaughlin & P. A. Lemke. Berlin:<br />

Springer-Verlag, pp. 57 83.<br />

Baura, G., Szaro, T. M. & Bruns, T. D. (1992). Gastrosuillus<br />

laricinus is a recent derivative of Suillus grevillei:<br />

molecular data. Mycologia, 84, 592 597.<br />

Bayman, P. & Cotty, J. P. (1993). Genetic diversity in<br />

Aspergillus flavus: association with afla<strong>to</strong>xin production<br />

and morphology. Canadian Journal of Botany, 71,<br />

23 31.<br />

Bayne, H. G. & Michener, H. D. (1979). Heat resistance of<br />

Byssochlamys ascospores. Applied and Environmental<br />

Microbiology, 37, 449 453.<br />

Beadle, G. W. & Tatum, E. L. (1941). Genetic control of<br />

biochemical reactions in Neurospora. Proceedings of the<br />

National Academy of Sciences of the USA, 27, 499 506.<br />

Beakes, G. W. (1981). Ultrastructural aspects of oospore<br />

differentiation. In The Fungal Spore: Morphogenetic<br />

Controls, ed. H. R. Hohl & G. Turian. London:<br />

Academic Press, pp. 71 94.<br />

Beakes, G. W. (1983). A comparative account of cyst coat<br />

on<strong>to</strong>geny in saprophytic and fish-lesion (pathogenic)<br />

isolates of the Saprolegnia diclina-parasitica complex.<br />

Canadian Journal of Botany, 61, 603 625.<br />

Beakes, G. W. (1987). Oomycete phylogeny: ultrastructural<br />

perspectives. In Evolutionary Biology of the <strong>Fungi</strong>,<br />

ed. A. D. M. Rayner, C. M. Brasier & D. Moore.<br />

Cambridge: Cambridge University Press,<br />

pp. 405 421.<br />

Beakes, G. W. (1998). Relationships between lower<br />

fungi and pro<strong>to</strong>zoa. In Evolutionary Relationships<br />

among Pro<strong>to</strong>zoa, ed. G. H. Coombs, K. Vickerman,<br />

M. A. Sleigh & A. Warren. London: Chapman & Hall,<br />

pp. 351 373.<br />

Beakes, G. W. & Gay, J. L. (1978a). Light and electron<br />

microscopy of oospore maturation in Saprolegnia<br />

furcata. 1. Cy<strong>to</strong>plasmic changes. Transactions of the<br />

British Mycological Society, 71, 11 24.<br />

Beakes, G. W. & Gay, J. L. (1978b). Light and electron<br />

microscopy of oospore maturation in Saprolegnia<br />

furcata. 2. Wall development. Transactions of the British<br />

Mycological Society, 71, 25 35.<br />

Beakes, G. W. & Glockling, S. L. (1998). Injection tube<br />

differentiation in gun cells of a Hap<strong>to</strong>glossa species<br />

which infects nema<strong>to</strong>des. Fungal Genetics and Biology,<br />

24, 45 68.


REFERENCES<br />

709<br />

Beakes, G. W. & Glockling, S. L. (2000). An ultrastructural<br />

analysis of organelle arrangement during gun<br />

(infection) cell differentiation in the nema<strong>to</strong>de<br />

parasite Hap<strong>to</strong>glossa dickii. Mycological Research, 104,<br />

1258 1269.<br />

Beakes, G. W., Canter, H. M. & Jaworski, G. H. M. (1993).<br />

Sporangium differentiation and zoospore finestructure<br />

of the chytrid Rhizophydium plank<strong>to</strong>nicum, a<br />

fungal parasite of Asterionella formosa. Mycological<br />

Research, 97, 1059 1074.<br />

Beaumont, A. (1947). Dependence on the weather of<br />

the date of pota<strong>to</strong> blight epidemics. Transactions of the<br />

British Mycological Society, 31, 45 53.<br />

Bécard, G. & Fortin, J. A. (1988). Early events of<br />

vesicular arbuscular mycorrhiza formation in Ri T-<br />

DNA transformed roots. New Phy<strong>to</strong>logist, 108, 211 218.<br />

Beckett, A. (1976a). Ultrastructural studies on exogenously<br />

dormant ascospores of Daldinia concentrica.<br />

Canadian Journal of Botany, 54, 689 697.<br />

Beckett, A. (1976b). Ultrastructural studies on germinating<br />

ascospores of Daldinia concentrica. Canadian<br />

Journal of Botany, 54, 698 705.<br />

Beckett, A. (1979a). Ultrastructure and development of<br />

the ascospore germ slit in Xylaria longipes. Transactions<br />

of the British Mycological Society, 72, 269 276.<br />

Beckett, A. (1979b). Comparative ultrastructure of the<br />

ascospore germ slit in Xylaria and Daldinia.<br />

Transactions of the British Mycological Society, 72,<br />

320 322.<br />

Beckett, A. (1981). The ascus with an apical pore:<br />

development, composition and function. In<br />

Ascomycete Systematics, the Luttrellian Concept, ed.<br />

D. R. Reynolds. New York: Springer-Verlag, pp. 7 26.<br />

Beckett, A. & Crawford, R. M. (1973). The development<br />

and fine structure of the ascus apex and its role<br />

during spore discharge in Xylaria longipes. New<br />

Phy<strong>to</strong>logist, 72, 357 369.<br />

Beckett, A. & Wilson, I. M. (1968). Ascus cy<strong>to</strong>logy of<br />

Podospora anserina. Journal of General Microbiology, 53,<br />

81 87.<br />

Beckett, A., Bar<strong>to</strong>n, R. & Wilson, I. M. (1968). Fine<br />

structure of the wall and appendage formation in<br />

ascospores of Podospora anserina. Journal of General<br />

Microbiology, 53, 89 94.<br />

Beckett, A., Heath, I. B. & McLaughlin, D. J. (1974). An<br />

Atlas of Fungal Ultrastructure. London: Longman.<br />

Beckman, C. H. (1987). The Nature of Wilt Diseases of<br />

Plants. St. Paul, USA: APS Press.<br />

Beever, R. E. & Weeds, P. L. (2004). Taxonomy and genetic<br />

variation of Botrytis and Botryotinia.InBotrytis: Biology,<br />

Pathology and Control, ed. Y. Elad, B. Williamson, P.<br />

Tudzynski & N. Delen. Dordrecht: Kluwer, pp. 29 52.<br />

Beever, R. E., Redgewell, R. J. & Dempsey, G. P. (1978).<br />

Function of rodlets on the surface of fungal spores.<br />

Nature, 272, 608 610.<br />

Begerow, D., Bauer, R. & Oberwinkler, F. (1997).<br />

Phylogenetic studies on large subunit ribosomal<br />

DNA sequences of smut fungi and related taxa.<br />

Canadian Journal of Botany, 75, 2045 2056.<br />

Begerow, D., Bauer, R. & Boekhout, T. (2000).<br />

Phylogenetic placements of<br />

ustilaginomyce<strong>to</strong>us anamorphs as deduced from<br />

nuclear LSU rDNA sequences. Mycological Research,<br />

104, 53 60.<br />

Begerow, D., Bauer, R. & Oberwinkler, F. (2002). The<br />

Exobasidiales: an evolutionary hypothesis.<br />

Mycological Progress, 1, 187 199.<br />

Bélanger, R. R. & Labbé, C. (2002). Control of powdery<br />

mildews without chemicals: prophylactic and biological<br />

alternatives for horticultural crops. In The<br />

Powdery Mildews. A Comprehensive Treatise, ed. R. R.<br />

Bélanger, W. R. Bushnell, A. J. Dik & T. L. W. Carver. St.<br />

Paul, USA: APS Press, pp. 256 267.<br />

Bell, A. & Mahoney, D. P. (1995). Coprophilous fungi in<br />

New Zealand. I. Podospora species with swollen<br />

agglutinated perithecial hairs. Mycologia, 87,<br />

375 396.<br />

Bell, A. & Mahoney, D. P. (1996). Perithecium development<br />

in Podospora tetraspora and Podospora vesticola.<br />

Mycologia, 88, 163 170.<br />

Bell, A. & Mahoney, D. P. (1997). Coprophilous fungi in<br />

New Zealand. II. Podospora species with coriaceous<br />

perithecia. Mycologia, 89, 908 915.<br />

Bell, A. A. & Wheeler, M. H. (1986). Biosynthesis and<br />

functions of fungal melanins. Annual Review of<br />

Phy<strong>to</strong>pathology, 24, 411 451.<br />

Bellemère, A. (1994). Asci and ascospores in ascomycete<br />

systematics. In Ascomycete Systematics. Problems and<br />

Perspectives in the Nineties, ed. D. L. Hawksworth. New<br />

York: Plenum Press, pp. 111 126.<br />

Benjamin, R. K. (1962). A new Basidiobolus that forms<br />

microspores. Aliso, 5, 223 233.<br />

Benjamin, R. K. (1966). The merosporangium. Mycologia,<br />

58, 1 42.<br />

Benjamin, R. K. (1979). Zygomycetes and their spores. In<br />

The Whole Fungus 2, ed. B. Kendrick. Ottawa: National<br />

Museum of Natural Sciences; Kananaskis<br />

Foundation, pp. 573 616.<br />

Benjamin, R. K. & Mehrotra, B. S. (1963). Obligate<br />

azygospore formation in two species of Mucor<br />

(Mucorales). Aliso, 5, 235 245.<br />

Bennett, R. J. & Johnson, A. D. (2005). Mating in Candida<br />

albicans and the search for a sexual cycle. Annual<br />

Review of Microbiology, 59, 233 255.


710 REFERENCES<br />

Benny, G. L. (1992). Observations on the Thamnidiaceae<br />

(Mucorales). V. Thamnidium. Mycologia, 84, 834 842.<br />

Benny, G. L. (2001). Zygomycota: Trichomycetes. In The<br />

Mycota VIIA: Systematics and Evolution, ed. D. J.<br />

McLaughlin, E. G. McLaughlin & P. A. Lemke. Berlin:<br />

Springer-Verlag, pp. 147 160.<br />

Benny, G. L. & Benjamin, R. K. (1976). Observations on<br />

Thamnidiaceae (Mucorales). II. Chae<strong>to</strong>cladium,<br />

Cokeromyces, Mycotypha and Phascolomyces. Aliso, 8,<br />

391 424.<br />

Benny, G. L. & Benjamin, R. K. (1993). Observations on<br />

Thamnidiaceae. 6. Two new species of<br />

Dicho<strong>to</strong>mocladium and the zygospores of D. hesseltinei<br />

(Chae<strong>to</strong>cladiaceae). Mycologia, 85, 660 671.<br />

Benny, G. L., Humber, R. A., & Mor<strong>to</strong>n, J. B. (2001).<br />

Zygomycota: Zygomycetes. In The Mycota VIIA:<br />

Systematics and Evolution, ed. D. J. McLaughlin, E. G.<br />

McLaughlin & P. A. Lemke. Berlin: Springer-Verlag,<br />

pp. 113 146.<br />

Berbee, M. L. (1996). Loculoascomycete origins and<br />

evolution of filamen<strong>to</strong>us ascomycete morphology<br />

based on 18S rRNA gene sequence data. Molecular<br />

Biology and Evolution, 13, 462 470.<br />

Berbee, M. L. (2001). The phylogeny of plant and animal<br />

pathogens in the Ascomycota. Physiological and<br />

Molecular Plant Pathology, 59, 165 187.<br />

Berbee, M. L. & Taylor, J. W. (1992a). Two ascomycete<br />

classes based on fruiting-body characters and ribosomal<br />

DNA sequence. Molecular Biology and Evolution,<br />

9, 278 284.<br />

Berbee, M. L. & Taylor, J. W. (1992b). 18S ribosomal RNA<br />

gene sequence characters place the human pathogen<br />

Sporothrix schenckii in the genus Ophios<strong>to</strong>ma.<br />

Experimental Mycology, 16, 87 91.<br />

Berbee, M. L. & Taylor, J. W. (1993). Dating the evolutionary<br />

radiations of the true fungi. Canadian Journal<br />

of Botany, 71, 1114 1127.<br />

Berbee, M. L. & Taylor, J. W. (1999). Fungal phylogeny. In<br />

Molecular Fungal Biology, ed. R. P. Oliver & M.<br />

Schweizer. Cambridge: Cambridge University Press,<br />

pp. 21 77.<br />

Berbee, M. L. & Taylor, J. W. (2001). Fungal molecular<br />

evolution: gene trees and geologic time. In The Mycota<br />

VIIB: Systematics and Evolution, ed. D. J. McLaughlin,<br />

E. G. McLaughlin & P. A. Lemke. Berlin: Springer-<br />

Verlag, pp. 229 245.<br />

Berbee, M. L. & Wells, K. (1988). Ultrastructural studies<br />

of mi<strong>to</strong>sis and the septal pore apparatus in Tremella<br />

globospora. Mycologia, 80, 479 492.<br />

Berbee, M. L. & Wells, K. (1989). Light and electron<br />

microscopic studies of meiosis and basidium on<strong>to</strong>geny<br />

in Clavicorona pyxidata. Mycologia, 81, 20 41.<br />

Berbee, M. L., Yoshimura, A., Sugiyama, J. & Taylor, J. W.<br />

(1995). Is Penicillium monophyletic? An evaluation of<br />

phylogeny in the family Trichocomaceae from 18S,<br />

5.8S and ITS ribosomal DNA sequence data.<br />

Mycologia, 87, 210 222.<br />

Berbee, M. L., Pirseyedi, M. & Hubbard, S. (1999).<br />

Cochliobolus phylogenetics and the origin of known,<br />

highly virulent pathogens, inferred from ITS and<br />

glyceraldehyde-3-phosphate dehydrogenase gene<br />

sequences. Mycologia, 91, 964 977.<br />

Berbee, M. L., Payne, B. P., Zhang, G., Roberts, R. G. &<br />

Turgeon, B. G. (2003). Shared ITS DNA substitutions<br />

in isolates of opposite mating type reveal a recombining<br />

his<strong>to</strong>ry for three presumed asexual species in<br />

the filamen<strong>to</strong>us ascomyce<strong>to</strong>us genus Alternaria.<br />

Mycological Research, 107, 169 182.<br />

Berch, S. M., Allen, T. R. & Berbee, M. L. (2002).<br />

Molecular detection, community structure and<br />

phylogeny of ericoid mycorrhizal fungi. Plant and<br />

Soil, 244, 55 66.<br />

Bérdy, J. (1986). Further antibiotics with practical<br />

application. In Biotechnology 4: Microbial Products II, ed.<br />

H. Pape & H.-J. Rehm. Weinheim, Germany: VCH,<br />

pp. 465 507.<br />

Bergman, K., Burke, P. V., Cerdá-Olmedo, E., et al.<br />

(1969). Phycomyces. Bacteriological Reviews, 33,<br />

99 157.<br />

Berrie, A. D. (1975). Detritus, micro-organisms and<br />

animals in fresh water. In The Role of Terrestrial and<br />

Aquatic Organisms in Decomposition Processes, ed. J. M.<br />

Anderson & A. MacFadyen. Oxford: Blackwell<br />

Scientific, pp. 323 338.<br />

Berta, G. & Fusconi, A. (1983). Ascosporogenesis in Tuber<br />

magnatum. Transactions of the British Mycological Society,<br />

80, 201 207.<br />

Bertaux, J., Schmid, M., Prevost-Bourne, N. C., et al.<br />

(2003). In situ identification of intracellular bacteria<br />

related <strong>to</strong> Paenibacillus spp. in the mycelium of the<br />

ec<strong>to</strong>mycorrhizal fungus Laccaria bicolor S238N.<br />

Applied and Environmental Microbiology, 69,<br />

4243 4248.<br />

Bertrand, H. (2000). Role of mi<strong>to</strong>chondrial DNA in the<br />

senescence and hypovirulence of fungi and potential<br />

for plant disease control. Annual Review of<br />

Phy<strong>to</strong>pathology, 38, 397 422.<br />

Besl, H. & Bresinsky, A. (1997). Chemosystematics<br />

of Suillaceae and Gomphidiaceae (suborder Suillineae).<br />

Plant Systematics and Evolution, 206, 223 242.<br />

Bettiol, W. (1999). Effectiveness of cow’s milk against<br />

zucchini squash powdery mildew (Sphaerotheca fuliginea)<br />

in greenhouse conditions. Crop Protection, 18,<br />

489 492.


REFERENCES<br />

711<br />

Beuchat, L. R. (1995). Indigenous fermented foods. In<br />

Biotechnology 9: Enzymes, Biomass, Food and Feed, ed. G.<br />

Reed & T. W. Nagodawithana. Weinheim, Germany:<br />

VCH, pp. 505 559.<br />

Beuchat, L. R. & Toledo, R. T. (1977). Behaviour of<br />

Byssochlamys nivea ascospores in fruit syrups.<br />

Transactions of the British Mycological Society, 68, 65 71.<br />

Beyrle, H. F., Smith, S. E., Peterson, R. L. & Franco,<br />

C. M. M. (1995). Colonization of Orchis morio pro<strong>to</strong>corms<br />

by a mycorrhizal fungus: effects of nitrogen<br />

nutrition and glyphosate in modifying the<br />

responses. Canadian Journal of Botany, 73, 1128 1140.<br />

Bhatnagar, D., Yu, J. & Ehrlich, K. C. (2002). Toxins of<br />

filamen<strong>to</strong>us fungi. Chemical Immunology, 81,<br />

167 206.<br />

Bidar<strong>to</strong>ndo, M. I., Burghardt, B., Gebauer, G., Bruns,<br />

T. D. & Read, D. J. (2004). Changing partners in the<br />

dark: iso<strong>to</strong>pic and molecular evidence of ec<strong>to</strong>mycorrhizal<br />

liaisons between forest orchids and trees.<br />

Proceedings of the Royal Society, B 271, 1799 1806.<br />

Biffen, R. H. (1905). Mendel’s laws of inheritance and<br />

wheat breeding. Journal of Agricultural Science, 1,4 48.<br />

Bigelow, D. M., Olsen, M. W. & Gilbertson, R. L. (2005).<br />

Labyrinthula terrestris sp. nov., a new pathogen of turf<br />

grass. Mycologia, 97, 185 190.<br />

Bilay, V. T. & Lelley, J. I. (1997). Growth of mycelium of<br />

Agaricus bisporus on biomass and conidium of<br />

Humicola insolens. Journal of Applied Botany<br />

Angewandte Botanik, 71, 21 23.<br />

Bimbó, A., Liu, J. & Balasubramanian, M. K. (2005). Roles<br />

of Pdk1p, a fission yeast protein related <strong>to</strong> phosphoinositide-dependent<br />

protein kinase, in the regulation<br />

of mi<strong>to</strong>sis and cy<strong>to</strong>kinesis. Molecular Biology of<br />

the Cell, 16, 3162 3175.<br />

Binder, M. & Bresinsky, A. (2002). Derivation of a<br />

polymorphic lineage of Gasteromycetes from bole<strong>to</strong>id<br />

ances<strong>to</strong>rs. Mycologia, 94, 85 98.<br />

Binder, M. & Hibbett, D. S. (2002). Higher-level phylogenetic<br />

relationships of Homobasidiomycetes<br />

(mushroom-forming fungi) inferred from four rDNA<br />

regions. Molecular Phylogenetics and Evolution, 22,<br />

76 90.<br />

Binder, M., Hibbett, D. S. & Moli<strong>to</strong>ris, H. P. (2001).<br />

Phylogenetic relationships of the marine gasteromycete<br />

Nia vibrissa. Mycologia, 93, 679 688.<br />

Bishop, C. D. & Cooper, R. M. (1983). An ultrastructural<br />

study of root invasion in three vascular wilt diseases.<br />

Physiological Plant Pathology, 22, 15 27.<br />

Bistis, G. (1956). Sexuality in Ascobolus stercorarius. I.<br />

Morphology of the ascogonium; plasmogamy;<br />

evidence for a sexual hormonal mechanism.<br />

American Journal of Botany, 43, 389 394.<br />

Bistis, G. (1957). Sexuality in Ascobolus stercorarius. II.<br />

Preliminary experiments on various aspects of the<br />

sexual process. American Journal of Botany, 44,<br />

436 443.<br />

Bistis, G. N. (1983). Evidence for diffusible mating-type<br />

specific trichogyne attractants in Neurospora crassa.<br />

Experimental Mycology, 7, 292 295.<br />

Bistis, G. N. (1998). Physiological heterothallism and<br />

sexuality in euascomycetes: a partial his<strong>to</strong>ry. Fungal<br />

Genetics and Biology, 23, 213 222.<br />

Bistis, G. & Raper, J. R. (1963). Heterothallism and<br />

sexuality in Ascobolus stercorarius. American Journal of<br />

Botany, 50, 880 891.<br />

Black, W. (1952). A genetical basis for the classification<br />

of strains of Phy<strong>to</strong>phthora infestans. Proceedings of the<br />

Royal Society of Edinburgh, B65,36 51.<br />

Blackwell, E. M. (1943). The life his<strong>to</strong>ry of Phy<strong>to</strong>phthora<br />

cac<strong>to</strong>rum (Leb. & Cohn) Schroet. Transactions of the<br />

British Mycological Society, 26, 71 89.<br />

Blackwell, M. (1994). Minute mycological mysteries: the<br />

influence of arthropods on the lives of fungi.<br />

Mycologia, 86, 1 17.<br />

Blackwell, M. & Malloch, D. (1989). Similarity of<br />

Amphoromorpha and secondary capilliconidia of<br />

Basidiobolus. Mycologia, 81, 735 741.<br />

Blackwell, W. H. & Powell, M. J. (1999). The nomenclatural<br />

propriety of Rhizophlyctis rosea. Mycotaxon, 70,<br />

213 217.<br />

Blakeslee, A. F. (1906). Zygospore germinations in the<br />

Mucorineae. Annales Mycologici, 4, 1 28.<br />

Bland, C. E. & Charles, T. M. (1972). Fine structure of<br />

Pilobolus: surface and wall structure. Mycologia, 64,<br />

774 785.<br />

Blan<strong>to</strong>n, R. L. (1990). Phylum Acrasea. In Handbook of<br />

Pro<strong>to</strong>ctista, ed. L. Margulis, J. O. Corliss, M. Melkonian<br />

& D. J. Chapman. Bos<strong>to</strong>n: Jones & Bartlett, pp. 75 87.<br />

Blanz, P. A. & Gottschalk, M. (1984). A comparison of 5S<br />

ribosomal RNA nucleotide sequences from smut<br />

fungi. Systematic and Applied Microbiology, 5, 518 526.<br />

Blodgett, D. J. (2001). Fescue <strong>to</strong>xicosis. The Veterinary<br />

Clinics of North America. Equine Practice, 17, 567 577.<br />

Bluemke, W. & Schrader, J. (2001). Integrated<br />

bioprocess for enhanced production of natural<br />

flavors and fragrances by Cera<strong>to</strong>cystis moniliformis.<br />

Biomolecular Engineering, 17, 137 142.<br />

Bly, J. E., Lawson, L. A., Dale, D. J., et al. (1992). Winter<br />

saprolegniosis in channel catfish. Diseases of Aquatic<br />

Organisms, 13, 155 164.<br />

Bobbitt, T. F. & Nordin, J. H. (1982). Production and<br />

composition of an exocellular nigeran-protein<br />

complex isolated from cultures of Aspergillus<br />

awamori. Journal of Bacteriology, 150, 365 376.


712 REFERENCES<br />

Bock, K. R. (1962). Control of coffee leaf rust in Kenya<br />

Colony. Transactions of the British Mycological Society, 45,<br />

301 313.<br />

Boddy, L. (1993). Saprotrophic cord-forming fungi:<br />

warfare strategies and other ecological aspects.<br />

Mycological Research, 97, 641 655.<br />

Boddy, L. & Rayner, A. D. M. (1983). Origins of decay in<br />

living deciduous trees: the role of moisture content<br />

and a re-appraisal of the expanded concept of tree<br />

decay. New Phy<strong>to</strong>logist, 94, 623 641.<br />

Boddy, L., Gibbon, O. M. & Grundy, M. A. (1985). Ecology<br />

of Daldinia concentrica: effect of abiotic variables on<br />

mycelial extension and interspecific interactions.<br />

Transactions of the British Mycological Society, 85,<br />

201 211.<br />

Boehm, E. W. A., Wenstrom, J. C., McLaughlin, D. J., et al.<br />

(1992). An ultrastructural pachytene karyotype for<br />

Puccinia graminis f. sp. tritici. Canadian Journal of<br />

Botany, 70, 401 413.<br />

Boerema, G. H., de Gruyter, J., Noordeloos, M. E. &<br />

Hamers, M. E. C. (2004). Phoma Identification Manual.<br />

Wallingford, UK: CABI Publishing.<br />

Bogus, M. I. & Scheller, K. (2002). Extraction of an<br />

insecticidal protein fraction from the parasitic<br />

fungus Conidiobolus coronatus (En<strong>to</strong>mophthorales).<br />

Acta Parasi<strong>to</strong>logica, 47, 66 72.<br />

Bogus, M. I. & Szczepanik, M. (2000). His<strong>to</strong>pathology of<br />

Conidiobolus coronatus (En<strong>to</strong>mophthorales) infection<br />

in Galleria mellonella (Lepidoptera) larvae. Acta<br />

Parasi<strong>to</strong>logica, 45, 48 54.<br />

Boidin, J. (1971). Nuclear behaviour in the mycelium<br />

and the evolution of the Basidiomycetes. In Evolution<br />

in the Higher Basidiomycetes, ed. R. H. Petersen.<br />

Knoxville: University of Tennessee Press,<br />

pp. 129 148.<br />

Boland, G. J. & Hall, R. (1994). Index of plant hosts of<br />

Sclerotinia sclerotiorum. Canadian Journal of Plant<br />

Pathology, 16, 93 108.<br />

Bölker, M. (2001). Ustilago maydis a valuable model<br />

system for the study of fungal dimorphism and<br />

virulence. Microbiology, 147, 1395 1401.<br />

Bol<strong>to</strong>n, M. D., Thomma, B. P. H. J. & Nelson, B. D. (2006).<br />

Sclerotinia sclerotiorum (Lib.) de Bary: biology and<br />

molecular traits of a cosmopolitan pathogen.<br />

Molecular Plant Pathology, 7, 1 16.<br />

Bonner, J. T. (1944). A descriptive study of the<br />

development of the slime mold Dictyostelium<br />

discoideum. American Journal of Botany, 31,<br />

175 182.<br />

Bonner, J. T. (1999). The his<strong>to</strong>ry of the cellular slime<br />

moulds as a ‘model system’ for developmental<br />

biology. Journal of Biosciences, 24, 7 12.<br />

Bonner, J. T. (2001). First Signals: The Evolution of<br />

Multicellular Development. Prince<strong>to</strong>n and Oxford:<br />

Prince<strong>to</strong>n University Press.<br />

Bonner, J. T., Kane, K. K. & Levey, R. H. (1956). Studies on<br />

the mechanics of growth in the common mushroom.<br />

Mycologia, 48, 13 19.<br />

Boone, C., Bussey, H., Greene, D., Thomas, D. Y. &<br />

Vernet, T. (1986). Yeast killer <strong>to</strong>xin: site-directed<br />

mutations implicate the precursor protein as the<br />

immunity component. Cell, 46, 105 113.<br />

Booth, C. (1971). The Genus Fusarium. Kew:<br />

Commonwealth Mycological Institute.<br />

Borel, J. F. (1986). Cyclosporin and its future. Progress in<br />

Allergy, 38, 9 18.<br />

Borges-Walmsley, M. I., Chen, D., Shu, X. & Walmsley,<br />

A. R. (2002). The pathobiology of Paracoccidioides<br />

brasiliensis. Trends in Microbiology, 10, 80 87.<br />

Borg-Karlson, A.-K., Englund, F. O. & Unelius, C. R.<br />

(1994). Dimethyl oligosulphides, major volatiles<br />

released from Sauromatum guttatum and Phallus<br />

impudicus. Phy<strong>to</strong>chemistry, 35, 321 323.<br />

Borneman, A. R., Hynes, M. J. & Andrianopoulos, A.<br />

(2000). The abaA homologue of Penicillium marneffei<br />

participates in two developmental programmes:<br />

conidiation and dimorphic growth. Molecular<br />

Microbiology, 38, 1034 1047.<br />

Borno, C. & van der Kamp, B. J. (1975). Timing of<br />

infection and development of Gymnosporangium<br />

fuscum on Juniperus. Canadian Journal of Botany, 53,<br />

1266 1269.<br />

Borriss, H. (1934). Beiträge zur Wachstums- und<br />

Entwicklungsphysiologie der Fruchtkörper von<br />

Coprinus lagopus. Planta, 22, 28 69.<br />

Bos, C. J. & Swart, K. (1995). Genetics of Aspergillus.InThe<br />

Mycota II: Genetics and Biotechnology, ed. U. Kück.<br />

Berlin: Springer-Verlag, pp. 19 33.<br />

Botella, M. A., Parker, J. E., Frost, L. N., et al. (1998).<br />

Three genes of the Arabidopsis RPP1 complex<br />

resistance locus recognize distinct Peronospora<br />

parasitica avirulence determinants. Plant Cell, 10,<br />

1847 1860.<br />

Boucias, D. G. & Pendland, J. C. (1998). Principles of<br />

Insect Pathology. Bos<strong>to</strong>n: Kluwer Academic<br />

Publishers.<br />

Bourett, T. M. & Howard, R. J. (1990). In vitro development<br />

of penetration structures in the rice blast<br />

fungus, Magnaporthe grisea. Canadian Journal of Botany,<br />

68, 329 342.<br />

Bourett, T. M. & Howard, R. J. (1992). Actin in<br />

penetration pegs of the fungal rice blast<br />

pathogen, Magnaporthe grisea. Pro<strong>to</strong>plasma,<br />

168, 20 26.


REFERENCES<br />

713<br />

Bourett, T. M., Czymmek, K. J. & Howard, R. J. (1998). An<br />

improved method for affinity probe localization in<br />

whole cells of filamen<strong>to</strong>us fungi. Fungal Genetics and<br />

Biology, 24, 3 13.<br />

Bourke, A. (1991). Pota<strong>to</strong> blight in Europe in 1845: the<br />

scientific controversy. In Phy<strong>to</strong>phthora, ed. J. A. Lucas,<br />

R. C. Shat<strong>to</strong>ck, D. S. Shaw & L. R. Cooke. Cambridge:<br />

Cambridge University Press, pp. 12 24.<br />

Bowden, J., Gregory, P. H. & Johnson, C. G. (1971).<br />

Possible wind transport of coffee leaf rust across the<br />

Atlantic Ocean. Nature, 229, 500 501.<br />

Bowman, B. H., Taylor, J. W., Brownlee, A. G., et al.<br />

(1992). Molecular evolution of the fungi: relationship<br />

of the Basidiomycetes, Ascomycetes and<br />

Chytridiomycetes. Molecular Biology and Evolution, 9,<br />

285 296.<br />

Boxma, B., Voncken, F., Jannink, S., et al. (2004). The<br />

anaerobic chytridiomycete fungus Piromyces sp. E2<br />

produces ethanol via pyruvate : formate lyase and an<br />

alcohol dehydrogenase E. Molecular Microbiology, 51,<br />

1389 1399.<br />

Bozarth, R. F., Koltin, Y., Weissman, M. B., et al. (1981).<br />

The molecular weight and packaging of dsRNAs in<br />

the mycovirus from Ustilago maydis killer strains.<br />

Virology, 113, 492 502.<br />

Boze, H., Moulin, G. & Galzy, P. (1995). Production of<br />

microbial biomass. In Biotechnology 9: Enzymes,<br />

Biomass, Food and Feed, ed. G. Reed &<br />

T. W. Nagodawithana. Weinheim, Germany: VCH,<br />

pp. 167 220.<br />

Bracker, C. E. (1968a). The ultrastructure and development<br />

of sporangia of Gilbertella persicaria. Mycologia,<br />

60, 1016 1067.<br />

Bracker, C. E. (1968b). Ultrastructure of the haus<strong>to</strong>rial<br />

apparatus of Erysiphe graminis and its relationship<br />

<strong>to</strong> the epidermal cell of barley. Phy<strong>to</strong>pathology, 58,<br />

12 30.<br />

Bradbury, E. M., Inglis, R. J., Matthews, H. R. & Langan,<br />

T. A. (1974). Molecular basis of control of mi<strong>to</strong>tic cell<br />

division in eukaryotes. Nature, 249, 553 556.<br />

Bradley, W. H. (1967). Two aquatic fungi (Chytridiales)<br />

of Eocene age from the Green River Formation of<br />

Wyoming. American Journal of Botany, 54, 577 582.<br />

Bradshaw, A. D. (1959). Population differentiation in<br />

Agrostis tenuis Sibth. II. The incidence and significance<br />

of infection by Epichloe typhina. New Phy<strong>to</strong>logist,<br />

58, 310 315.<br />

Bradshaw, J. E., Gemmell, D. J. & Wilson, R. N. (1997).<br />

Transfer of resistance <strong>to</strong> clubroot (Plasmodiophora<br />

brassicae) <strong>to</strong> swedes (Brassica napus L. var. napobrassica<br />

Peterm.) from B. rapa. Annals of Applied Biology, 130,<br />

337 348.<br />

Bradshaw, R. E., Bennett, J. W. & Peberdy, J. F. (1983).<br />

Parasexual analysis of Aspergillus parasiticus. Journal of<br />

General Microbiology, 129, 2117 2123.<br />

Brady, K. C., O’Kiely, P., Forristal, P. D. & Fuller, H.<br />

(2005). Schizophyllum commune on big-bale grass silage<br />

in Ireland. Mycologist, 19, 30 35.<br />

Brandhorst, T. T., Rooney, P. J., Sullivan, T. D. & Klein, B.<br />

(2002). Molecular genetic analysis of Blas<strong>to</strong>myces<br />

dermatitidis reveals new insights about pathogenic<br />

mechanisms. International Journal of Medical<br />

Microbiology, 292, 363 371.<br />

Brasel<strong>to</strong>n, J. P. (1995). Current status of the<br />

Plasmodiophorids. Critical Reviews in Microbiology, 21,<br />

263 275.<br />

Brasel<strong>to</strong>n, J. P. (2001). Plasmodiophoromycota. In The<br />

Mycota VIIA: Systematics and Evolution, ed. D. J.<br />

McLaughlin, E. G. McLaughlin & P. A. Lemke. Berlin:<br />

Springer-Verlag, pp. 81 91.<br />

Brasier, C. M. (1975a). Stimulation of sex organ formation<br />

in Phy<strong>to</strong>phthora by antagonistic species of<br />

Trichoderma. I. The effect in vitro. New Phy<strong>to</strong>logist, 74,<br />

183 194.<br />

Brasier, C. M. (1975b). Stimulation of sex organ formation<br />

in Phy<strong>to</strong>phthora by antagonistic species of<br />

Trichoderma. II. Ecological implications. New<br />

Phy<strong>to</strong>logist, 74, 195 198.<br />

Brasier, C. M. (1991a). Current questions in Phy<strong>to</strong>phthora<br />

systematics: the role of the population approach. In<br />

Phy<strong>to</strong>phthora, ed. J. A. Lucas, R. C. Shat<strong>to</strong>ck, D. S. Shaw<br />

& L. R. Cooke. Cambridge: Cambridge University<br />

Press, pp. 104 128.<br />

Brasier, C. M. (1991b). Ophios<strong>to</strong>ma novo-ulmi sp. nov.,<br />

causative agent of the current Dutch elm disease<br />

pandemics. Mycopathologia, 115, 151 161.<br />

Brasier, C. M. (1993). The genetic system as a fungal<br />

taxonomic <strong>to</strong>ol: gene flow, molecular variation and<br />

sibling species in the Ophios<strong>to</strong>ma piceae Ophios<strong>to</strong>ma<br />

ulmi complex and its taxonomic and ecological<br />

significance. In Cera<strong>to</strong>cystis and Ophios<strong>to</strong>ma. Taxonomy,<br />

Ecology and Pathogenicity, ed. M. J. Wingfield, K. A.<br />

Seifert & J. F. Webber. St. Paul, USA: APS Press,<br />

pp. 77 92.<br />

Brasier, C. M. (2001). Rapid evolution of introduced<br />

plant pathogens via interspecific hybridization.<br />

BioScience, 51, 123 133.<br />

Brasier, C. M. & Kirk, S. A. (2001). Designation of the<br />

EAN and NAN races of Ophios<strong>to</strong>ma novo-ulmi as<br />

subspecies. Mycological Research, 105, 547 554.<br />

Brasier, C. M. & Mehrotra, M. D. (1995). Ophios<strong>to</strong>ma himalulmi<br />

sp. nov., a new species of Dutch elm disease<br />

fungus endemic in the Himalayas. Mycological<br />

Research, 99, 205 215.


714 REFERENCES<br />

Brasier, C. M., Kirk, S. A., Pipe, N. D. & Buck, K. W. (1998).<br />

Rare interspecific hybrids in natural populations of<br />

the Dutch elm disease pathogens Opios<strong>to</strong>ma ulmi and<br />

O. novo-ulmi. Mycological Research, 102, 45 57.<br />

Brasier, C. M., Kirk, S. A., Delcan, J., et al. (2004).<br />

Phy<strong>to</strong>phthora alni sp. nov. and its variants: designation<br />

of emerging heteroploid hybrid pathogens spreading<br />

on Alnus trees. Mycological Research, 108, 1172 1184.<br />

Braun, U. (1987). A monograph of the Erysiphales<br />

(powdery mildews). Nova Hedwigia Beihefte, 89, 1 700.<br />

Braun, U. (1995). The Powdery Mildews (Erysiphales) of<br />

Europe. Jena: G. Fischer Verlag.<br />

Braun, U., Cook, R. T. A., Inman, A. J. & Shin, H.-D.<br />

(2002). The taxonomy of the powdery mildews. In The<br />

Powdery Mildews. A Comprehensive Treatise, ed. R. R.<br />

Bélanger, W. R. Bushnell, A. J. Dik, & T. L. W. Carver.<br />

St. Paul, USA: APS Press, pp. 13 55.<br />

Bravery, A. F. (1991). The strategy for the eradication of<br />

Serpula lacrymans. In Serpula lacrymans. Fundamental<br />

Biology and Control Strategies, ed. D. H. Jennings & A. F.<br />

Bravery. Chichester: John Wiley & Sons, pp. 117 130.<br />

Breeze, E. M. & Dix, N. J. (1981). Seasonal analysis of the<br />

fungal community on Acer platanoides leaves.<br />

Transactions of the British Mycological Society, 77,<br />

321 328.<br />

Brefeld, O. (1876). Ueber die Zygosporenbildung bei<br />

Mortierella rostafinskii nebst Bemerkungen über die<br />

Systematik der Zygomyceten. Sitzungsberichte der<br />

Gesellschaft Naturforschender Freunde zu Berlin, 91.<br />

Brefeld, O. (1881). Botanische Untersuchungen über<br />

Schimmelpilze. 4. Pilobolus. Untersuchungen aus dem<br />

Gesammtgebiete der Mykologie IV. Lepizig: Arthur Felix.<br />

Brefeld, O. (1888). Basidiomyceten II.<br />

Pro<strong>to</strong>basidiomyceten. Untersuchungen aus dem<br />

Gesammtgebiete der Mykologie VII. Leipzig: Arthur<br />

Felix.<br />

Breitenbach, J. & Kränzlin, F. (1984). <strong>Fungi</strong> of Switzerland.<br />

1. Ascomycetes. Lucerne: Verlag Mykologia.<br />

Breitenbach, M. & Simon-Nobbe, B. (2002). The allergens<br />

of Cladosporium herbarum and Alternaria alternata.<br />

Chemical Immunology, 81, 48 72.<br />

Bresinsky, A. & Besl, H. (1990). A Colour Atlas of Poisonous<br />

<strong>Fungi</strong>. London: Manson Publishing.<br />

Brewer, J. G. & Boerema, G. H. (1965). Electron microscope<br />

observations on the development of pycnidiospores<br />

in Phoma and Ascochyta spp. Proceedings of<br />

the Academy of Sciences, Amsterdam, C68,86 97.<br />

Brian, P. W. (1960). Griseofulvin. Transactions of the<br />

British Mycological Society, 43, 1 13.<br />

Bridgmon, G. H. (1959). Production of new races of<br />

Puccinia graminis var. tritici by vegetative fusion.<br />

Phy<strong>to</strong>pathology, 49, 386 388.<br />

Brierley, W. B. (1918). The microconidia of Botrytis<br />

cinerea. Kew Bulletin, 129 146.<br />

Brobyn, P. J. & Wilding, N. (1977). Invasive and developmental<br />

processes in En<strong>to</strong>mophthora species infecting<br />

aphids. Transactions of the British Mycological Society,<br />

69, 349 366.<br />

Brobyn, P. J. & Wilding, N. (1983). Invasive and developmental<br />

processes in En<strong>to</strong>mophthora muscae infecting<br />

houseflies. Transactions of the British Mycological<br />

Society, 80, 1 8.<br />

Brodie, H. J. (1936). The occurrence and function of<br />

oidia in the Hymenomycetes. American Journal of<br />

Botany, 23, 309 327.<br />

Brodie, H. J. (1975). The Bird’s Nest <strong>Fungi</strong>. Toron<strong>to</strong>:<br />

Toron<strong>to</strong> University Press.<br />

Brodie, H. J. (1984). More bird’s nest fungi. A supplement<br />

<strong>to</strong> ‘The Bird’s Nest <strong>Fungi</strong>’ (1975). Lejeunia N. S.,<br />

112, 1 69.<br />

Brodie, H. J. & Gregory, P. H. (1953). The action of<br />

wind in the dispersal of spores from cup-shaped<br />

plant structures. Canadian Journal of Botany, 31,<br />

402 410.<br />

Brodo, I. M. & Richardson, D. H. S. (1978). Chimeroid<br />

associations in the genus Peltigera. Lichenologist, 10,<br />

157 170.<br />

Brodo, I. M., Sharnoff, S. D. & Sharnoff, S. (2001). Lichens<br />

of North America. New Haven: Yale University Press.<br />

Brooke, A. G. (1994). Industrial fermentation and<br />

Aspergillus citric acid. In The Genus Aspergillus. From<br />

Taxonomy and Genetics <strong>to</strong> Industrial Application, ed. K. A.<br />

Powell, A. Renwick & J. F. Peberdy. New York: Plenum<br />

Press, pp. 129 134.<br />

Brown, A. J. & Cassel<strong>to</strong>n, L. A. (2001). Mating in mushrooms:<br />

increasing the chances but prolonging the<br />

affair. Trends in Genetics, 17, 393 400.<br />

Brown, A. J. P. & Gow, N. A. R. (2002). Signal transduction<br />

and morphogenesis in Candida albicans. InThe<br />

Mycota VIII: Biology of the Fungal Cell, ed. R. J. Howard<br />

& N. A. R. Gow. Berlin: Springer-Verlag, pp. 55 71.<br />

Brown, J. K. M. (2002). Comparative genetics of<br />

avirulence and fungicide resistance in the powdery<br />

mildew fungi. In The Powdery Mildews. A Comprehensive<br />

Treatise, ed. R. R. Bélanger, W. R. Bushnell, A. J. Dik &<br />

T. L. W. Carver. St. Paul, USA: APS Press, pp. 56 65.<br />

Brown, J. K. M. & Hovmøller, M. S. (2002). Aerial dispersal<br />

of pathogens on the global and continental scales<br />

and its impact on plant disease. Science, 297,<br />

537 541.<br />

Brown, J. K. M., Simpson, C. G. & Wolfe, M. S. (1993).<br />

Adaptation of barley powdery mildew populations in<br />

England <strong>to</strong> varieties with two resistance genes. Plant<br />

Pathology, 42, 108 115.


REFERENCES<br />

715<br />

Brown, J. S., Whan, J. H., Kenny, M. K. & Merriman, P. R.<br />

(1995). The effect of coffee leaf rust on foliation and<br />

yield of coffee in Papua New Guinea. Crop Protection,<br />

14, 589 592.<br />

Bruggink, A., Straathof, A. J. & van der Wielen, L. A.<br />

(2003). A ‘fine’ chemical industry for life science<br />

products: green solutions <strong>to</strong> chemical challenges.<br />

Advances in Biochemical Engineering/Biotechnology, 80,<br />

69 113.<br />

Bruin, G. C. A. & Edging<strong>to</strong>n, L. V. (1983). The chemical<br />

control of diseases caused by zoosporic fungi. A<br />

many-sided problem. In Zoosporic Plant Pathogens,<br />

ed. S. T. Buczacki. London: Academic Press,<br />

pp. 193 232.<br />

Bruning, K. (1991a). Infection of the dia<strong>to</strong>m Asterionella<br />

by a chytrid. I. Effects of light on reproduction and<br />

infectivity of the parasite. Journal of Plank<strong>to</strong>n Research,<br />

13, 103 117.<br />

Bruning, K. (1991b). Effects of phosphorus limitation<br />

on the epidemiology of a chytrid phy<strong>to</strong>plank<strong>to</strong>n<br />

parasite. Freshwater Biology, 25, 409 417.<br />

Bruning, K. (1991c). Effects of temperature and light on<br />

the population dynamics of the<br />

Asterionella Rhizophydium association. Journal of<br />

Plank<strong>to</strong>n Research, 13, 707 719.<br />

Brunner, D. & Nurse, P. (2000). New concepts in fission<br />

yeast morphogenesis. Philosophical Transactions of the<br />

Royal Society, B 355, 873 877.<br />

Bruns, T. D., White, T. J. & Taylor, J. W. (1991). Fungal<br />

molecular systematics. Annual Review of Ecology and<br />

Systematics, 22, 525 564.<br />

Bruns, T. D., Vilgalys, R., Barns, S. M., et al. (1992).<br />

Evolutionary relationships within the fungi:<br />

analyses of nuclear small subunit rRNA<br />

sequences. Molecular Phylogenetics and Evolution,<br />

1, 231 241.<br />

Bruns, T. D., Szaro, T. M., Gardes, M., et al. (1998). A<br />

sequence database for the identification of ec<strong>to</strong>mycorrhizal<br />

basidiomycetes by phylogenetic analysis.<br />

Molecular Ecology, 7, 257 272.<br />

Bryan, G. T., Daniels, M. J. & Osbourn, A. E. (1995).<br />

Comparison of fungi within the<br />

Gaeumannomyces Phialophora complex by analysis of<br />

ribosomal DNA sequences. Applied and Environmental<br />

Microbiology, 61, 681 689.<br />

Buchanan, K. L. & Murphy, J. W. (1998). What makes<br />

Cryp<strong>to</strong>coccus neoformans a pathogen? Emerging<br />

Infectious Diseases, 4, 71 83.<br />

Buchta, M. & Cvak, L. (1999). Ergot alkaloids and other<br />

metabolites of the genus Claviceps. In Ergot. The Genus<br />

Claviceps, ed. V. Křen & L. Cvak. Amsterdam: Harwood<br />

Academic Publishers, pp. 173 200.<br />

Buczacki, S. T. (1983). Plasmodiophora, an interrelationship<br />

between biological and practical<br />

problems. In Zoosporic Plant Pathogens: A Modern<br />

Perspective, ed. S. T. Buczacki. London: Academic Press,<br />

pp. 161 191.<br />

Buczacki, S. T., Toxopeus, H., Mattusch, P., et al. (1975).<br />

Study of physiologic specialization in Plasmodiophora<br />

brassicae: proposals for an attempted rationalization<br />

through an international approach. Transactions of<br />

the British Mycological Society, 65, 295 303.<br />

Büdel, B. & Scheidegger, C. (1996). Thallus morphology<br />

and ana<strong>to</strong>my. In Lichen Biology, ed. T. H. Nash.<br />

Cambridge: Cambridge University Press, pp. 37 64.<br />

Budge, S. P. & Whipps, J. M. (2001). Potential for<br />

integrated control of Sclerotinia sclerotiorum in glasshouse<br />

lettuce using Coniothyrium minitans and<br />

reduced fungicide application. Phy<strong>to</strong>pathology, 91,<br />

221 227.<br />

Buller, A. H. R. (1909). Researches on <strong>Fungi</strong>, 1. London:<br />

Longmans Green & Co.<br />

Buller, A. H. R. (1922). Researches on <strong>Fungi</strong>, 2. London:<br />

Longmans Green & Co.<br />

Buller, A. H. R. (1924). Researches on <strong>Fungi</strong>, 3. London:<br />

Longmans Green & Co.<br />

Buller, A. H. R. (1931). Researches on <strong>Fungi</strong>, 4. London:<br />

Longmans, Green & Co.<br />

Buller, A. H. R. (1933). Researches on <strong>Fungi</strong>, 5. London:<br />

Longmans, Green & Co.<br />

Buller, A. H. R. (1934). Researches on <strong>Fungi</strong>, 6. London:<br />

Longmans, Green & Co.<br />

Buller, A. H. R. (1950). Researches on <strong>Fungi</strong>, 7. Toron<strong>to</strong>:<br />

Royal Society of Canada; Toron<strong>to</strong> University Press.<br />

Buller, A. H. R. & Vanterpool, T. C. (1933). The violent<br />

discharge of the basidiospores (secondary conidia) of<br />

Tilletia tritici. In A. H. R. Buller, Researches on <strong>Fungi</strong>, 5,<br />

pp. 207 278. London: Longmans, Green & Co.<br />

Bullerwell, C. E. & Lang, B. F. (2005). Fungal evolution:<br />

the case of the vanishing mi<strong>to</strong>chondrion. Current<br />

Opinion in Microbiology, 8, 362 369.<br />

Bulmer, G. S. (1964). Spore germination of forty-two<br />

species of puffballs. Mycologia, 56, 630 632.<br />

Bultman, T. L., White, J. F., Bodish, T. I. & Welch, A. M.<br />

(1998). A new kind of mutualism by insects and<br />

fungi. Mycological Research, 102, 235 238.<br />

Bunyard, B. A., Nicholson, M. S. & Royse, D. F. (1995).<br />

Phylogenetic resolution of Morchella, Verpa, and<br />

Disciotis (Pezizales: Morchellaceae) based on restriction<br />

enzyme analysis of the 28S ribosomal RNA gene.<br />

Experimental Mycology, 19, 223 233.<br />

Burdsall, H. H. (1981). The taxonomy of Sporotrichum<br />

pruinosum and Sporotrichum pulverulentum/<br />

Phanerochaete chrysosporium. Mycologia, 73, 675 680.


716 REFERENCES<br />

Burdsall, H. H. & Banik, M. T. (2001). The genus<br />

Laetiporus in North America. Harvard Papers in Botany,<br />

6, 43 55.<br />

Burgeff, H. (1920). Über den Parasitismus des<br />

Chae<strong>to</strong>cladium und die heterokaroyotische Natur der<br />

von ihm auf Mucorineenerzeugten Gallen. Zeitschrift<br />

für Botanik, 12, 1 35.<br />

Burgeff, H. (1924). Untersuchungen über Sexualität<br />

und Parasitismus bei Mucorineen. I. Botanische<br />

Abhandlungen, 4, 1 135.<br />

Burnett, J. H. (1965). The natural his<strong>to</strong>ry of recombination<br />

systems. In Incompatibilty in <strong>Fungi</strong>, ed. K. Esser &<br />

J. R. Raper. Berlin: Springer-Verlag, pp. 98 113.<br />

Burnett, J. H. (1976). Fundamentals of Mycology, 2nd edn.<br />

London: Edward Arnold.<br />

Burnett, J. H. & Boulter, M. E. (1963). The mating<br />

systems of fungi. II. Mating systems of the<br />

Gasteromycetes Mycocalia denudata and M. duriaeana.<br />

New Phy<strong>to</strong>logist, 62, 217 236.<br />

Burney, D. A., Robinson, G. S. & Pigott Burney, L. (2003).<br />

Sporormiella and the late Holocene extinctions in<br />

Madagascar. Proceedings of the National Academy of<br />

Sciences of the USA, 100, 10 800 10 8005.<br />

Burr, A. W. & Beakes, G. W. (1994). Characterization of<br />

zoospore and cyst surface structure in saprophytic<br />

and fish pathogenic Saprolegnia species (oomycete<br />

fungal protists). Pro<strong>to</strong>plasma, 181, 142 163.<br />

Burt, A., Carter, D. A., Koenig, G. L., White, T. J. & Taylor,<br />

J. W. (1996). Molecular markers reveal cryptic sex in<br />

the human pathogen Coccidioides immitis. Proceedings<br />

of the National Academy of Sciences of the USA, 93,<br />

770 773.<br />

Buscot, F. (1989). Field observations on growth and<br />

development of Morchella rotunda and Mitrophora semilibera<br />

in relation <strong>to</strong> forest soil temperature. Canadian<br />

Journal of Botany, 67, 589 593.<br />

Buscot, F. (1994). Ec<strong>to</strong>mycorrhizal types and endobacteria<br />

associated with ec<strong>to</strong>mycorrhizas of<br />

Morchella elata (Fr.) Boudier with Picea abies (L.)<br />

Karst. Mycorrhiza, 4, 223 232.<br />

Buscot, F. & Kottke, I. (1990). His<strong>to</strong>cy<strong>to</strong>logical study of the<br />

association between Morchella esculenta and the root<br />

system of Picea abies. New Phy<strong>to</strong>logist, 116, 425 430.<br />

Buscot, F. & Roux, J. (1987). Association between living<br />

roots and ascocarps of Morchella rotunda (Pers.)<br />

Boudier. Transactions of the British Mycological Society,<br />

89, 249 252.<br />

Bushnell, W. R. & Gay, J. (1978). Accumulation of<br />

solutes in relation <strong>to</strong> the structure and function of<br />

haus<strong>to</strong>ria in powdery mildews. In The Powdery<br />

Mildews, ed. D. M. Spencer. London, New York and<br />

San Francisco: Academic Press, pp. 183 235.<br />

Bushnell, W. R. & Roelfs, A. P., eds. (1984). The Cereal<br />

Rusts 1: Origins, Specificity, Structure, and Physiology.<br />

Orlando: Academic Press.<br />

Butler, E. J. & Jones, S. G. (1949). Plant Pathology. London:<br />

Macmillan.<br />

Butler, G. M. (1966). Vegetative structure. In The <strong>Fungi</strong>.<br />

An Advanced Treatise II: The Fungal Organism. New York<br />

and London: Academic Press, pp. 83 112.<br />

Butler, M. J. & Day, A. W. (1998). Fungal melanins: a<br />

review. Canadian Journal of Microbiology, 44,<br />

1115 1136.<br />

Butt, T. M., Beckett, A. & Wilding, N. (1981). Pro<strong>to</strong>plasts<br />

of the in vivo life cycle of Erynia neoaphidis. Journal of<br />

General Microbiology, 127, 417 421.<br />

Butt, T. M., Beckett, A. & Wilding, N. (1990). A<br />

his<strong>to</strong>logical study of the invasive and developmental<br />

process of the aphid pathogen Erynia neoaphidis<br />

(Zygomycotina: En<strong>to</strong>mophthorales) in the pea aphid<br />

Acyrthosiphon pisum. Canadian Journal of Botany, 68,<br />

2153 2163.<br />

Butterfield, N. J. (2005). Probable Proterozoic fungi.<br />

Paleobiology, 31, 165 182.<br />

Byrde, R. J. W. & Willetts, H. J. (1977). The Brown Rot <strong>Fungi</strong><br />

of Fruit. Their Biology and Control. Oxford: Pergamon<br />

Press.<br />

Cai, L., Jeewon, R. & Hyde, K. D. (2005). Phylogenetic<br />

evaluation and taxonomic revision of Schizothecium<br />

based on ribosomal DNA and protein coding genes.<br />

Fungal Diversity, 19, 1 21.<br />

Cairney, J. W. G. (1991). Rhizomorphs: organs of<br />

exploration or exploitation? Mycologist, 5, 5 10.<br />

Cairney, J. W. G. (1992). Translocation of solutes in<br />

ec<strong>to</strong>mycorrhizal and saprotrophic rhizomorphs.<br />

Mycological Research, 96, 135 141.<br />

Cairney, J. W. G. (2000). Evolution of mycorrhiza<br />

systems. Naturwissenschaften, 76, 467 475.<br />

Cairney, J. W. G. (2002). Pisolithus death of the<br />

pan-global super fungus. New Phy<strong>to</strong>logist, 153,<br />

199 201.<br />

Cairney, J. W. G. & Chambers, S. M. (1997). Interactions<br />

between Pisolithus tinc<strong>to</strong>rius and its hosts: a review of<br />

current knowledge. Mycorrhiza, 7, 117 131.<br />

Çakar, Z. P., Sauer, U., Bailey, J. E., et al. (2000). Vacuolar<br />

morphology and cell cycle distribution are modified<br />

by leucine limitation in auxotrophic Saccharomyces<br />

cerevisiae. Biologie Cellulaire, 92, 629 637.<br />

Calderoni, M., Sieber, T. N. & Holdenrieder, O. (2003).<br />

Stereum sanguinolentum: is it an amphithallic basidiomycete?<br />

Mycologia, 95, 232 238.<br />

Calduch, M., Gené, J., Cano, J., Stchigel, A. M. & Guarro,<br />

J. (2004). Three new species of Oidiodendron Robak<br />

from Spain. Studies in Mycology, 50, 159 170.


REFERENCES<br />

717<br />

Callac, P. (1995). Breeding of edible fungi with<br />

emphasis among the French genetic resources of<br />

Agaricus bisporus. Canadian Journal of Botany, 73,<br />

S980 S986.<br />

Callac, P., Jacobé de Haut, I., Imbernon, M., et al. (2003).<br />

A novel homothallic variety of Agaricus bisporus<br />

comprises rare tetrasporic isolates from Europe.<br />

Mycologia, 95, 222 231.<br />

Callaghan, A. A. (1969a). Light and spore discharge in<br />

En<strong>to</strong>mophthorales. Transactions of the British<br />

Mycological Society, 53, 87 97.<br />

Callaghan, A. A. (1969b). Morphogenesis in Basidiobolus<br />

ranarum. Transactions of the British Mycological Society,<br />

53, 99 108.<br />

Callaghan, A. A. (2004). Teaching techniques for mycology:<br />

22. Disclosure of Conidiobolus and Basidiobolus<br />

from soils and leaf litters. Mycologist, 18, 29 32.<br />

Callan, B. E. (1993). Rhizina root rot of conifers. Forest<br />

Pest Leaflet, 56. Vic<strong>to</strong>ria, B. C.: Pacific Forestry Centre.<br />

Callot, G. (1999). La Truffe, la Terre, et la Vie. Paris: INRA.<br />

Calvo-Mendez, C. & Ruiz-Herrera, J. (1987). Biosynthesis<br />

of chi<strong>to</strong>san in membrane fractions from Mucor rouxii<br />

by the concerted action of chitin synthetase and a<br />

particulate deacetylase. Experimental Mycology, 11,<br />

128 140.<br />

Câmara, M. P. S., Palm, M. E., van Berkum, P. & O’Neill,<br />

N. R. (2002). Molecular phylogeny of Lep<strong>to</strong>sphaeria and<br />

Phaeosphaeria. Mycologia, 94, 630 640.<br />

Campbell, A. H. (1933). Zone lines in plant tissues I. The<br />

black lines formed by Xylaria polymorpha (Pers.) Grey<br />

in hardwoods. Annals of Applied Biology, 20, 123 145.<br />

Campbell, C. L. & Long, D. L. (2001). The campaign <strong>to</strong><br />

eradicate the common barberry in the United States.<br />

In Stem Rust of Wheat. From Ancient Enemy <strong>to</strong> Modern<br />

Foe, ed. P. D. Peterson. St. Paul, USA: APS Press,<br />

pp. 16 50.<br />

Campbell, G. F. & Crous, P. W. (2003). Genetic stability<br />

of net spot hybrid progeny of the barley pathogen<br />

Pyrenophora teres. Australasian Plant Pathology, 32,<br />

283 287.<br />

Campbell, R. (1968). An electron microscope study of<br />

spore structure and development in Alternaria brassicicola.<br />

Journal of General Microbiology, 54, 381 392.<br />

Campbell, R. N. (1985). Longevity of Olpidium brassicae in<br />

air-dry soil and the persistence of the lettuce big-vein<br />

agent. Canadian Journal of Botany, 63, 2288 2289.<br />

Campbell, R. N. (1996). Fungal transmission of plant<br />

viruses. Annual Review of Phy<strong>to</strong>pathology, 34, 87 108.<br />

Campbell, R. N. & Fry, P. R. (1966). The nature of the<br />

associations between Olpidium brassicae and lettuce<br />

big vein and <strong>to</strong>bacco necrosis viruses. Virology, 29,<br />

222 233.<br />

Canham, S. C. (1969). Taxonomy and morphology of<br />

Hypocrea citrina. Mycologia, 61, 315 331.<br />

Cannon, P. F. & Minter, D. W. (1986). The<br />

Rhytismataceae of the Indian subcontinent.<br />

Mycological Papers, 155, 1 123.<br />

Canter, H. M. (1950). Studies on British chytrids. IX.<br />

Anisolpidium stigeoclonii (de Wildeman) n. comb.<br />

Transactions of the British Mycological Society, 33,<br />

335 344.<br />

Canter, H. M. & Jaworski, G. H. M. (1978). The isolation,<br />

maintenance and host range studies of the chytrid<br />

Rhizophydium plank<strong>to</strong>nicum Canter emend., parasitic<br />

on Asterionella formosa Hassall. Annals of Botany N. S.,<br />

42, 967 979.<br />

Canter, H. M. & Jaworski, G. H. M. (1979). The occurrence<br />

of a hypersensitive reaction in the plank<strong>to</strong>nic<br />

dia<strong>to</strong>m Asterionella formosa Hassall parasitized by the<br />

chytrid Rhizophydium plank<strong>to</strong>nicum Canter emend., in<br />

culture. New Phy<strong>to</strong>logist, 82, 187 206.<br />

Canter, H. M. & Jaworski, G. H. M. (1981). The effect of<br />

light and darkness upon infection of Asterionella<br />

formosa Hassall by the chytrid Rhizophydium<br />

plank<strong>to</strong>nicum Canter emend. Annals of Botany N.<br />

S., 47, 13 30.<br />

Canter, H. M. & Lund, J. W. G. (1948). Studies on<br />

plank<strong>to</strong>n parasites. I. Fluctuations in the numbers of<br />

Asterionella formosa Hass., in relation <strong>to</strong> fungal<br />

epidemics. New Phy<strong>to</strong>logist, 47, 238 261.<br />

Canter, H. M. & Lund, J. W. G. (1953). The parasitism of<br />

dia<strong>to</strong>ms with special reference <strong>to</strong> lakes in the<br />

English Lake District. Transactions of the British<br />

Mycological Society, 36, 13 37.<br />

Canter, H. M. & Willoughby, L. G. (1964). A parasitic<br />

Blas<strong>to</strong>cladiella from Windermere plank<strong>to</strong>n. Journal of<br />

the Royal Microscopical Society, 83, 365 372.<br />

Cantino, E. C. (1955). Physiology and phylogeny in the<br />

water-molds a re-evaluation. Quarterly Review of<br />

Biology, 25, 269 277.<br />

Cantino, E. C., Lovett, J. S., Leak, L. V. & Lythgoe, J.<br />

(1963). The single mi<strong>to</strong>chondrion, fine structure,<br />

and germination of the spore of Blas<strong>to</strong>cladiella<br />

emersonii. Journal of General Microbiology, 31,<br />

393 404.<br />

Cantino, E. C., Truesdell, L. C. & Shaw, D. S. (1968). Life<br />

his<strong>to</strong>ry of the motile spore of Blas<strong>to</strong>cladiella emersonii:<br />

a study in cell differentiation. Journal of the Elisha<br />

Mitchell Scientific Society, 84, 125 146.<br />

Cao, H., Li, X. & Dong, X. (1998). Generation of broadspectrum<br />

disease resistance by overexpression of an<br />

essential regula<strong>to</strong>ry gene in systemic acquired<br />

resistance. Proceedings of the National Academy of<br />

Sciences of the USA, 95, 6531 6536.


718 REFERENCES<br />

Carbone, I. & Kohn, L. M. (1993). Ribosomal DNA<br />

sequence divergence within internal transcribed<br />

spacer 1 of the Sclerotiniaceae. Mycologia, 85,<br />

415 427.<br />

Carisse, O., Philion, V., Rolland, D. & Bernier, J. (2000).<br />

Effect of fall application of fungal antagonists on<br />

spring ascospore production of the apple scap<br />

pathogen, Venturia inaequalis. Phy<strong>to</strong>pathology, 90,<br />

31 37.<br />

Carlile, M. J. (1971). Myxomycetes and other slime<br />

moulds. In Methods in Microbiology, ed. C. Booth. New<br />

York: Academic Press, pp. 237 265.<br />

Carlile, M. J. (1983). Motility, taxis, and tropism in<br />

Phy<strong>to</strong>phthora. InPhy<strong>to</strong>phthora. Its Biology, Taxonomy,<br />

Ecology, and Pathology, ed. D. C. Erwin, S. Bartnicki-<br />

Garcia & P. H. Tsao. St. Paul: APS Press, pp. 95 107.<br />

Carlile, M. J. (1995). The success of the hypha and<br />

mycelium. In The Growing Fungus, ed. N. A. R. Gow &<br />

G. M. Gadd. London: Chapman & Hall, pp. 3 19.<br />

Carlile, M. J. (1996a). The discovery of fungal sex<br />

hormones: I. Sirenin. Mycologist, 10, 3 6.<br />

Carlile, M. J. (1996b). The discovery of fungal sex<br />

hormones: II. Antheridiol. Mycologist, 10, 113 117.<br />

Carlile, M. J. & Dee, J. (1967). Plasmodial and lethal<br />

interaction between strains in a myxomycete.<br />

Nature, 215, 832 834.<br />

Caron, C. (1995). Commercial production of baker’s<br />

yeast and wine yeast. In Biotechnology 9: Enzymes,<br />

Biomass, Food and Feed, ed. G. Reed & T. W.<br />

Nagodawithana. Weinheim, Germany: VCH,<br />

pp. 321 351.<br />

Carpenter, S. E. & Trappe, J. M. (1985). Phoenicoid fungi:<br />

a proposed term for fungi that fruit after heat<br />

treatment of substrates. Mycotaxon, 23, 203 206.<br />

Carreiro, M. M. & Koske, R. E. (1992). Room temperature<br />

isolations can bias against selection of low<br />

temperature microfungi in temperate forest soils.<br />

Mycologia, 84, 886 900.<br />

Carroll, F. E. & Carroll, G. C. (1971). Fine structural<br />

studies on ‘poroconidium’ formation in Stemphylium<br />

botryosum. In Taxonomy of <strong>Fungi</strong> Imperfecti, ed. W. B.<br />

Kendrick. Toron<strong>to</strong>: Toron<strong>to</strong> University Press,<br />

pp. 75 91.<br />

Carruthers, R. I., Haynes, D. L. & MacLeod, D. M. (1985).<br />

En<strong>to</strong>mophthora muscae (En<strong>to</strong>mophthorales:<br />

En<strong>to</strong>mophthoraceae) mycosis in the onion fly Delia<br />

antiqua (Diptera: Anthomyiidae). Journal of<br />

Invertebrate Pathology, 45, 81 93.<br />

Carver, T. L. W. & Bushnell, W. R. (1983). The probable<br />

role of primary germ tubes in water uptake before<br />

infection by Erysiphe graminis. Physiological and<br />

Molecular Plant Pathology, 31, 237 250.<br />

Carver, T. L. W. & Ingerson, S. M. (1987). Responses of<br />

Erysiphe graminis germlings <strong>to</strong> contact with artificial<br />

and host surfaces. Physiological and Molecular Plant<br />

Pathology, 30, 359 372.<br />

Carver, T. L. W., Ingerson-Morris, S. M., Thomas, B. J. &<br />

Zeyen, R. J. (1995). Early interactions during powdery<br />

mildew infection. Canadian Journal of Botany, 73,<br />

S632 S639.<br />

Carver, T. L. W., Kunoh, H., Thomas, B. J. & Nicholson,<br />

R. L. (1999). Release and visualization of the extracellular<br />

matrix of conidia of Blumeria graminis.<br />

Mycological Research, 103, 547 560.<br />

Casadevall, A. & Perfect, J. R. (1998). Cryp<strong>to</strong>coccus neoformans.<br />

Washing<strong>to</strong>n, DC: ASM Press.<br />

Cassel<strong>to</strong>n, L. A. (2002). Mate recognition in fungi.<br />

Heredity, 88, 142 147.<br />

Cassel<strong>to</strong>n, L. A. & Economou, A. (1985). Dikaryon<br />

formation. In Developmental Biology of Higher <strong>Fungi</strong>, ed.<br />

D. Moore, L. A. Cassel<strong>to</strong>n, D. A. Wood, & J. C.<br />

Frankland. Cambridge: Cambridge University Press,<br />

pp. 213 229.<br />

Cassel<strong>to</strong>n, L. A. & Olesnicky, N. S. (1998). Molecular<br />

genetics of mating recognition in basidiomycete<br />

fungi. Microbiological and Molecular Biology Reviews, 62,<br />

55 70.<br />

Cassel<strong>to</strong>n, L. A. & Riquelme, M. (2004). Genetics of<br />

Coprinus. In The Mycota II: Genetics and Biotechnology,<br />

2nd edn, ed. U. Kück. Berlin: Springer-Verlag,<br />

pp. 37 52.<br />

Castle, A., Speranzini, D., Rghei, N., et al. (1998).<br />

Morphological and molecular identification of<br />

Trichoderma isolates on North American mushroom<br />

farms. Applied and Environmental Microbiology, 64,<br />

133 137.<br />

Castle, A. J. & Day, A. W. (1984). Isolation and identification<br />

of a-<strong>to</strong>copherol as an inducer of the parasitic<br />

phase of Ustilago violacea. Phy<strong>to</strong>pathology, 74,<br />

1194 1200.<br />

Castle, A. J., S<strong>to</strong>cco, N. & Boulianne, R. (1996). Fimbrialdependent<br />

mating in Microbotryum violaceum involves<br />

a mannose lectin interaction. Canadian Journal of<br />

Microbiology, 42, 461 466.<br />

Castle, E. S. (1942). Spiral growth and reversal of<br />

spiraling in Phycomyces, and their bearing on primary<br />

wall structure. American Journal of Botany, 29,<br />

664 672.<br />

Castlebury, L. A. & Domier, L. L. (1998). Small subunit<br />

ribosomal RNA gene phylogeny of Plasmodiophora<br />

brassicae. Mycologia, 90, 102 107.<br />

Castlebury, L. A. & Glawe, D. A. (1993). A comparison of<br />

three techniques for inoculating Chinese cabbage<br />

with Plasmodiophora brassicae. Mycologia, 85, 866 867.


REFERENCES<br />

719<br />

Castlebury, L. A., Rossman, A. Y., Jaklitsch, W. J. &<br />

Vasilyeva, L. N. (2002). A preliminary overview of the<br />

Diaporthales based on large subunit nuclear ribosomal<br />

DNA sequences. Mycologia, 94, 1017 1031.<br />

Castro, C. D., Koretsky, A. P. & Domach, M. M. (1999).<br />

NMR-observed phosphate trafficking and polyphosphate<br />

dynamics in wild-type and vph1 1 mutant<br />

Saccharomyces cerevisiae in response <strong>to</strong> stresses.<br />

Biotechnology Progress, 15, 65 73.<br />

Catanzariti, A.-M., Dodds, P. N., Lawrence, G. J., Ayliffe,<br />

M. A. & Ellis, J. G. (2006). Haus<strong>to</strong>rially expressed<br />

secreted proteins from flax rust are highly enriched<br />

for avirulence elici<strong>to</strong>rs. Plant Cell, 18, 243 256.<br />

Caten, C. E. & Jinks, J. E. (1966). Heterokaryosis: its<br />

significance in wild homothallic ascomycetes and<br />

<strong>Fungi</strong> Imperfecti. Transactions of the British Mycological<br />

Society, 49, 81 93.<br />

Cavalier-Smith, T. (1981). Eukaryote kingdoms: seven or<br />

nine? BioSystems, 14, 461 481.<br />

Cavalier-Smith, T. (1986). The kingdom Chromista:<br />

origin and systematics. In Progress in Phycological<br />

Research, ed. F. E. Round & D. J. Chapman. Bris<strong>to</strong>l:<br />

Biopress, pp. 309 47.<br />

Cavalier-Smith, T. (1987). The origin of fungi and<br />

pseudofungi. In Evolutionary Biology of <strong>Fungi</strong>, ed.<br />

A. D. M. Rayner, C. M. Brasier & D. Moore. Cambridge:<br />

Cambridge University Press, pp. 339 53.<br />

Cavalier-Smith, T. (1998). A revised six-kingdom system<br />

for life. Biological Reviews, 73, 203 266.<br />

Cavalier-Smith, T. (2001). What are fungi? In The Mycota<br />

VIIA: Systematics and Evolution, ed. D. J. McLaughlin,<br />

E. G. McLaughlin & P. A. Lemke. Berlin: Springer-<br />

Verlag, pp. 3 37.<br />

Cavender, J. C. (1990). Phylum Dictyostelida. In<br />

Handbook of Pro<strong>to</strong>ctista, ed. L. Margulis, J. O. Corliss, M.<br />

Melkonian & D. J. Chapman. Bos<strong>to</strong>n: Jones & Bartlett,<br />

pp. 88 101.<br />

Cecca<strong>to</strong>-An<strong>to</strong>nini, S. R. & Sudbery, P. E. (2004).<br />

Filamen<strong>to</strong>us growth in Saccharomyces cerevisiae.<br />

Brazilian Journal of Microbiology, 35, 173 181.<br />

Celerin, M. & Day, A. W. (1998). Sex, smut, and RNA: the<br />

complexity of fungal fimbriae. International Journal of<br />

Plant Sciences, 159, 175 184.<br />

Celerin, M., Laudenbach, D. E., Bancroft, J. B. & Day,<br />

A. W. (1994). Evidence that fimbriae of the smut<br />

fungus Microbotryum violaceum contain RNA.<br />

Microbiology, 140, 2699 2704.<br />

Celerin, M., Ray, J. M., Schisler, N. J., et al. (1996). Fungal<br />

fimbriae are composed of collagen. EMBO Journal, 15,<br />

4445 4453.<br />

Cerdá-Olmedo, E. (1975). The genetics of Phycomyces<br />

blakesleeanus. Genetical Research, 25, 285 296.<br />

Cerdá-Olmedo, E. (2001). Phycomyces and the biology of<br />

light and color. FEMS Microbiology Reviews, 25,<br />

503 512.<br />

Cerdá-Olmedo, E. & Lipson, E. D., eds. (1987). Phycomyces.<br />

New York: Cold Spring Harbor Labora<strong>to</strong>ry Press.<br />

Cerenius, L., Söderhäll, K., Persson, M. & Ajaxon, R.<br />

(1988). The crayfish plague fungus, Aphanomyces<br />

astaci diagnosis, isolation and pathology.<br />

Freshwater Crayfish, 7, 131 144.<br />

Certik, M. & Shimizu, S. (1999). Biosynthesis and<br />

regulation of microbial polyunsaturated fatty acid<br />

production. Journal of Bioscience and Bioengineering, 87,<br />

1 14.<br />

Cessna, S. G., Sears, V. E., Dickman, M. B. & Low, P. S.<br />

(2000). Oxalic acid, a pathogenicity fac<strong>to</strong>r for<br />

Sclerotinia sclerotiorum, suppresses the oxidative burst<br />

of the host plant. Plant Cell, 12, 2192 2199.<br />

Chadefaud, M. (1942). Études d’asques. II. Structure et<br />

ana<strong>to</strong>mie comparée de l’appareil apical des asques<br />

chez divers Discomycètes et Pyrenomycètes. Revue de<br />

Mycologie, 7, 57 88.<br />

Chalutz, E. & Wilson, C. L. (1990). Postharvest biocontrol<br />

of green and blue mold and sour rot of citrus<br />

fruit by Debaryomyces hansenii. Plant Disease, 74,<br />

134 137.<br />

Chambers, S. M. & Cairney, J. W. G. (1999). Pisolithus. In<br />

Ec<strong>to</strong>mycorrhizal <strong>Fungi</strong>: Key Genera in Profile, ed. J. W. G.<br />

Cairney & S. M. Chambers. Berlin: Springer-Verlag,<br />

pp. 1 31.<br />

Chambers, T. C. & Willoughby, L. G. (1964). The fine<br />

structure of Rhizophlyctis rosea, a soil<br />

Phycomycete. Journal of the Royal Microscopical Society,<br />

83, 355 364.<br />

Chambers, T. C., Markus, K. & Willoughby, L. G. (1967).<br />

The fine structure of the mature zoosporangium of<br />

Nowakowskiella profusa. Journal of General Microbiology,<br />

46, 135 141.<br />

Chamier, A. C. (1985). Cell-wall degrading enzymes of<br />

aquatic hyphomycetes: a review. Botanical Journal of<br />

the Linnean Society, 91, 67 81.<br />

Champavier, Y., Pommier, M.-T., Aprin, N., Voiland, A.<br />

& Pellon, G. (2000). 10-oxo-trans-8-decenoic acid<br />

(ODA); production, biological activities, and<br />

comparison of other hormone-like substances in<br />

Agaricus bisporus. Enzyme and Microbial Technology,<br />

26, 243 251.<br />

Chang, C. W., Yang, H. C. & Leu, L. S. (1984). Zygospore<br />

formation of Choanephora cucurbitarum. Transactions of<br />

the Mycological Society of Japan, 25, 67 74.<br />

Chang, S. T. & Miles, P. G. (2004). Mushrooms: Cultivation,<br />

Nutritional Value, Medicinal Effect and Environmental<br />

Impact, 2nd edn. Boca Ra<strong>to</strong>n: CRC Press.


720 REFERENCES<br />

Chang, S. T. & Yau, C. K. (1971). Volvariella volvacea and its<br />

life his<strong>to</strong>ry. American Journal of Botany, 58, 552 561.<br />

Chapela, I. H. (1989). <strong>Fungi</strong> in healthy stems and<br />

branches of American beech and aspen: a comparative<br />

study. New Phy<strong>to</strong>logist, 113, 65 75.<br />

Chapela, I. H. & Boddy, L. (1988a). Fungal colonization<br />

of attached beech branches. I. Early stages of<br />

development of fungal communities. New Phy<strong>to</strong>logist,<br />

110, 39 45.<br />

Chapela, I. H. & Boddy, L. (1988b). Fungal colonization<br />

of attached beech branches. II. Spatial and temporal<br />

organization of communities arising from latent<br />

invaders in bark and functional sapwood under<br />

different moisture regimes. New Phy<strong>to</strong>logist, 110,<br />

47 57.<br />

Chapela, I. H., Petrini, O. & Petrini, L. E. (1990). Unusual<br />

ascospore germination in Hypoxylon fragiforme: first<br />

steps in the establishment of an endophytic<br />

symbiosis. Canadian Journal of Botany, 68, 2571 2575.<br />

Chapela, I. H., Petrini, O. & Hagmann, L. (1991).<br />

Monolignol glucosides as specific recognition<br />

messengers in fungal/plant symbioses. Physiological<br />

and Molecular Plant Pathology, 39, 289 298.<br />

Chapela, I. H., Petrini, O. & Bielser, G. (1993). The<br />

physiology of ascospore eclosion in Hypoxylon fragiforme:<br />

mechanisms in the early recognition and<br />

establishment of an endophytic symbiosis.<br />

Mycological Research, 97, 157 162.<br />

Charnley, A. K. & St. Leger, R. J. (1991). The role of<br />

cuticle-degrading enzymes in fungal pathogenesis in<br />

insects. In The Fungal Spore and Disease Initiation in<br />

Plants and Animals, ed. G. T. Cole & H. C. Hoch. New<br />

York: Plenum Press, pp. 267 286.<br />

Chaubal, R., Wilmot, V. A. & Wynn, W. K. (1991).<br />

Visualization, adhesiveness, and cy<strong>to</strong>chemistry of<br />

the extracellular matrix produced by germ tubes of<br />

Puccinia sorghi. Canadian Journal of Botany, 69,<br />

2044 2054.<br />

Chaverri, P. & Samuels, G. J. (2003). Hypocrea/Trichoderma<br />

(Ascomycota, Hypocreales, Hypocreaceae): species with<br />

green ascospores. Studies in Mycology, 48, 1 116.<br />

Chaverri, P., Samuels, G. J. & Stewart, E. L. (2001).<br />

Hypocrea virens, the teleomorph of Trichoderma virens.<br />

Mycologia, 93, 1113 1124.<br />

Chazalet, V., Debeaupuis, J.-P., Sarfati, J., et al. (1998).<br />

Molecular typing of environmental and patient<br />

isolates of Aspergillus fumigatus from various hospital<br />

settings. Journal of Clinical Microbiology, 36,<br />

1494 1500.<br />

Chelkowski, J. & Visconti, A., eds. (1992). Alternaria:<br />

Biology, Plant Diseases and Metabolites. Amsterdam:<br />

Elsevier.<br />

Chen, B., Gao, S., Choi, G. H. & Nuss, D. L. (1996).<br />

Extensive alteration of fungal gene transcript accumulation<br />

and elevation of G-protein-regulated cAMP<br />

levels by a virulence-attenuating hypovirus.<br />

Proceedings of the National Academy of Sciences of the USA,<br />

93, 7996 8000.<br />

Chen, C.-J. (1998). Morphological and molecular studies<br />

in the genus Tremella. Bibliotheca Mycologica, 174,<br />

1 225.<br />

Chen, T.-H., Hsu, C.-S., Tsai, P.-J., Ho, Y.-F. & Lin, N.-S.<br />

(2001). Heterotrimeric G-protein and signal transduction<br />

in the nema<strong>to</strong>de-trapping fungus<br />

Arthrobotrys dactyloides. Planta, 212, 858 863.<br />

Chen, X. M. (2005). Epidemiology and control of<br />

stripe rust (Puccinia striiformis f. sp. tritici) on<br />

wheat. Canadian Journal of Plant Pathology, 27,<br />

314 337.<br />

Cheng, S. & Tu, C. C. (1978). Auricularia spp. In The<br />

Biology and Cultivation of Edible Mushrooms, ed. S. T.<br />

Chang & W. A. Hayes. New York: Academic Press,<br />

pp. 605 625.<br />

Cherniak, R. & Sundstrom, J. B. (1994). Polysaccharide<br />

antigens of the capsule of Cryp<strong>to</strong>coccus neoformans.<br />

Infection and Immunity, 62, 1507 1512.<br />

Cherry, J. M., Ball, C., Weng, S., et al. (1997). Genetic and<br />

physical maps of Saccharomyces cerevisiae. Nature, 387,<br />

67 73.<br />

Chesters, C. G. C. & Greenhalgh, G. N. (1964).<br />

Geniculosporium serpens gen. et sp. nov., the imperfect<br />

state of Hypoxylon serpens. Transactions of the British<br />

Mycological Society, 47, 393 401.<br />

Chet, I. (1987). Trichoderma application, mode of<br />

action, and potential as a biocontrol agent of<br />

soilborne plant pathogenic fungi. In Innovative<br />

Approaches <strong>to</strong> Plant Disease Control, ed. I. Chet. New<br />

York: John Wiley & Sons, pp. 137 160.<br />

Chet, I., Henis, Y. & Kislev, N. (1969). Ultrastructure of<br />

sclerotia and hyphae of Sclerotium rolfsii Sacc. Journal<br />

of General Microbiology, 57, 143 147.<br />

Chil<strong>to</strong>n, W. S. (1994). The chemistry and mode of action<br />

of mushroom <strong>to</strong>xins. In Handbook of Mushroom<br />

Poisoning. Diagnosis and Treatment, 2nd edn, ed. D. G.<br />

Spoerke & B. H. Rumack. Boca Ra<strong>to</strong>n: CRC Press,<br />

pp. 165 231.<br />

Chilvers, G. A., Lapeyrie, F. F. & Douglass, P. A. (1985).<br />

A contrast between Oomycetes and other taxa of<br />

mycelial fungi in regard <strong>to</strong> the metachromatic<br />

granule formation. New Phy<strong>to</strong>logist, 99, 203 210.<br />

Chiu, S. W. (1993). Evidence for a haploid life-cycle in<br />

Volvaria volvacea from microspectropho<strong>to</strong>metric<br />

measurements and observations of nuclear behaviour.<br />

Mycological Research, 97, 1481 1485.


REFERENCES<br />

721<br />

Chiu, S. W. & Moore, D. (1990a). A mechanism for gill<br />

pattern formation in Coprinus cinereus. Mycological<br />

Research, 94, 320 326.<br />

Chiu, S. W. & Moore, D. (1990b). Development of the<br />

basidiome of Volvariella bombycina. Mycological<br />

Research, 94, 327 337.<br />

Chiu, S. W., Moore, D. & Chang, S. T. (1989). Basidiome<br />

polymorphism in Volvariella bombycina. Mycological<br />

Research, 92, 69 77.<br />

Chong, J., Leonard, K. J. & Salmeron, J. J. (2000).<br />

A North American system of nomenclature for<br />

Puccinia coronata f. sp. avenae. Plant Disease, 84,<br />

580 585.<br />

Chou, H. H. & Wu, W. S. (2002). Phylogenetic analysis<br />

of internal transcribed spacer regions of the genus<br />

Alternaria, and the significance of<br />

filament-beaked conidia. Mycological Research,<br />

106, 164 169.<br />

Christensen, M. J. (1996). Antifungal activity in grasses<br />

infected with Acremonium and Epichloe endophytes.<br />

Australasian Plant Pathology, 25, 186 191.<br />

Christensen, M. J., Bennett, R. J. & Schmid, J.<br />

(2001). Vascular bundle colonization by<br />

Neotyphodium endophytes in natural and novel<br />

associations with grasses. Mycological Research,<br />

105, 1239 1245.<br />

Christensen, M. J., Bennett, R. J. & Schmid, J. (2002).<br />

Growth of Epichloe/Neotyphodium and p-endophytes in<br />

leaves of Lolium and Festuca grasses. Mycological<br />

Research, 106, 93 106.<br />

Christiansen, S. K., Wirsel, S., Yun, S.-H., Yoder, O. C. &<br />

Turgeon, B. G. (1998). The two Cochliobolus mating<br />

type genes are conserved among species but one of<br />

them is missing in C. vic<strong>to</strong>riae. Mycological Research,<br />

102, 919 929.<br />

Chung, K.-R. & Schardl, C. L. (1997a). Sexual cycle and<br />

horizontal transmission of the grass symbiont,<br />

Epichloe typhina. Mycological Research, 101, 295 301.<br />

Chung, K.-R. & Schardl, C. L. (1997b). Vegetative<br />

compatibility between and within Epichloe species.<br />

Mycologia, 89, 558 565.<br />

Cisar, C. R. & TeBeest, D. O. (1999). Mating system of the<br />

filamen<strong>to</strong>us ascomycete, Glomerella cingulata. Current<br />

Genetics, 35, 127 133.<br />

Clark, J., Haskins, E. F. & Stephenson, S. L. (2004).<br />

Culture and reproductive systems of 11 species of<br />

Myce<strong>to</strong>zoans. Mycologia, 96, 36 40.<br />

Clark, J. S. C. & Spencer-Phillips, P. T. N. (1993).<br />

Accumulation of pho<strong>to</strong>assimilate by Peronospora<br />

viciae (Berk.) Casp. and leaves of Pisum sativum L.:<br />

evidence for nutrient uptake via intercellular<br />

hyphae. New Phy<strong>to</strong>logist, 124, 107 119.<br />

Clarke, D. D. & Akhkha, A. (2002). Population genetics<br />

of powdery mildew-natural plant pathosystems. In<br />

The Powdery Mildews. A Comprehensive Treatise, ed. R. R.<br />

Bélanger, W. R. Bushnell, A. J. Dik & T. L. W. Carver. St.<br />

Paul, USA: APS Press, pp. 200 218.<br />

Claus, R., Hoppen, H. O. & Karg, H. (1981). The secret of<br />

truffles: a steroidal pheromone? Experientia, 37,<br />

1178 1179.<br />

Clausen, M., Kräuter, R., Schachermayr, G., Potrykus, I.<br />

& Sautter, C. (2000). Antifungal activity of a virally<br />

encoded gene in transgenic wheat. Nature<br />

Biotechnology, 18, 446 449.<br />

Clax<strong>to</strong>n, J. R., Potter, U. J., Blakesley, D. & Clarkson, J. M.<br />

(1996). An ultrastructural study of the interaction<br />

between Spongospora subterranea f. sp. nasturtii and<br />

watercress roots. Mycological Research, 100,<br />

1431 1439.<br />

Clay, C. M. & Walsh, J. A. (1997). Spongospora subterranea<br />

f. sp. nasturtii, ultrastructure of the plasmodial-host<br />

interface, food vacuoles, flagellar apparatus and exit<br />

pores. Mycological Research, 101, 737 744.<br />

Clay, K. (1988). Fungal endophytes of grasses: a<br />

defensive mutualism between plants and fungi.<br />

Ecology, 69, 10 16.<br />

Clay, K. & Schardl, C. (2002). Evolutionary origins<br />

and ecological consequences of endophyte<br />

symbiosis with grasses. American Naturalist, 160,<br />

S99 S127.<br />

Clémençon, H. (2004). Cy<strong>to</strong>logy and Plec<strong>to</strong>logy<br />

of the Hymenomycetes. Berlin and Stuttgart:<br />

J. Cramer.<br />

Clement, J. A., Porter, R., Butt, T. M. & Beckett, A. (1997).<br />

Characteristics of adhesion pads formed during<br />

imbibition and germination of urediniospores of<br />

Uromyces viciae-fabae on host and synthetic surfaces.<br />

Mycological Research, 101, 1445 1458.<br />

Clifford, B. C. (1985). Barley leaf rust. In The Cereal Rusts<br />

2: Diseases, Distribution, Epidemiology, and Control, ed.<br />

A. P. Roelfs & R. Bushnell. Orlando: Academic Press,<br />

pp. 173 205.<br />

Clutterbuck, A. J. (1995). Genetics of fungi. In The<br />

Growing Fungus, ed. N. A. R. Gow & G. M. Gadd.<br />

London: Chapman & Hall, pp. 239 253.<br />

Coates, D. & Rayner, A. D. M. (1985a).<br />

Heterokaryon homokaryon interactions in Stereum<br />

hirsutum. Transactions of the British Mycological Society,<br />

84, 637 645.<br />

Coates, D. & Rayner, A. D. M. (1985b). Genetic control<br />

and variation in expression of the ‘bow-tie’ reactions<br />

between homokaryons of Stereum hirsutum.<br />

Transactions of the British Mycological Society, 84,<br />

191 205.


722 REFERENCES<br />

Coates, D., Rayner, A. D. M. & Todd, N. K. (1981). Mating<br />

behaviour, mycelial antagonism and the establishment<br />

of individuals in Stereum hirsutum. Transactions<br />

of the British Mycological Society, 76, 41 51.<br />

Cocchiet<strong>to</strong>, M., Skert, N., Nimis, P. L. & Sava, G. (2002). A<br />

review on usnic acid, an interesting natural<br />

compound. Naturwissenschaften, 89, 137 146.<br />

Cochrane, B. J., Brown, J., Wain, R. P., Yangco, B. G. &<br />

Testrake, D. (1989). Genetic studies in the genus<br />

Basidiobolus I. Isozyme variation among isolates of<br />

human and natural populations. Mycologia, 81,<br />

504 513.<br />

Coffey, M. D. (1975). Ultrastructural features of the<br />

haus<strong>to</strong>rial apparatus of the white blister fungus<br />

Albugo candida. Canadian Journal of Botany, 53,<br />

1285 1299.<br />

Coffey, M. D. & Gees, R. (1991). The cy<strong>to</strong>logy of<br />

development. In Phy<strong>to</strong>phthora infestans, the Cause of<br />

Late Blight of Pota<strong>to</strong>, ed. D. S. Ingram & P. H. Williams.<br />

London: Academic Press, pp. 31 51.<br />

Coffey, M. D. & Wilson, U. E. (1983). An ultrastructural<br />

study of the late-blight fungus Phy<strong>to</strong>phthora infestans<br />

and its interaction with the foliage of two pota<strong>to</strong><br />

cultivars possessing different levels of general (field)<br />

resistance. Canadian Journal of Botany, 61, 2669 2685.<br />

Cole, G. T. (1975). The thallic mode of conidiogenesis in<br />

the <strong>Fungi</strong> Imperfecti. Canadian Journal of Botany, 53,<br />

2983 3001.<br />

Cole, G. T. (1986). Models of cell differentiation in<br />

conidial fungi. Microbiological Reviews, 50, 95 132.<br />

Cole, G. T. & Kendrick, W. B. (1969a). Conidium on<strong>to</strong>geny<br />

in hyphomycetes. The annellophores of<br />

Scopulariopsis brevicaulis. Canadian Journal of Botany, 47,<br />

925 929.<br />

Cole, G. T. & Kendrick, W. B. (1969b). Conidium on<strong>to</strong>geny<br />

in hyphomycetes. The arthrospores of<br />

Oidiodendron and Geotrichum, and the endoarthrospores<br />

of Sporendonema. Canadian Journal of Botany, 47,<br />

1773 1780.<br />

Cole, G. T. & Samson, R. A. (1979). Patterns of Development<br />

in Conidial <strong>Fungi</strong>. London: Pitman.<br />

Cole, L., Orlovich, D. A. & Ashford, A. E. (1998). Structure,<br />

function, and motility of vacuoles in filamen<strong>to</strong>us<br />

fungi. Fungal Genetics and Biology, 24, 86 100.<br />

Coley-Smith, J. R. (1959). Studies of the biology of<br />

Sclerotium cepivorum Berk. III. Host range; persistence<br />

and viability of sclerotia. Annals of Applied Biology, 47,<br />

511 518.<br />

Coley-Smith, J. R. (1986). A comparison of flavor and<br />

odor compounds of onion, leek, garlic and Allium<br />

fistulosum in relation <strong>to</strong> germination of sclerotia of<br />

Sclerotium cepivorum. Plant Pathology, 35, 370 376.<br />

Coley-Smith, J. R. & Cooke, R. C. (1971). Survival and<br />

germination of fungal sclerotia. Annual Review of<br />

Phy<strong>to</strong>pathology, 9, 65 92.<br />

Coley-Smith, J. R. & King, J. E. (1969). The production by<br />

species of Allium of alkyl sulphides and their effect<br />

on germination of sclerotia of Sclerotium cepivorum<br />

Berk. Annals of Applied Biology, 64, 289 301.<br />

Coley-Smith, J. R. & Parfitt, D. (1986). Some effects of<br />

diallyl disulphide on sclerotia of Sclerotium cepivorum:<br />

possible novel control method for white rot disease<br />

of onions. Pesticide Science, 37, 587 594.<br />

Coley-Smith, J. R., Verhoeff, K. & Jarvis, W. R., eds.<br />

(1980). The Biology of Botrytis. London: Academic Press.<br />

Colgan, W. & Claridge, A. W. (2002). Mycorrhizal<br />

effectiveness of Rhizopogon spores recovered from<br />

faecal pellets of small forest-dwelling mammals.<br />

Mycological Research, 106, 314 320.<br />

Collinge, A. J. & Trinci, A. P. J. (1974). Hyphal tips of wild<br />

type and spreading colonial mutants of Neurospora<br />

crassa. Archives of Microbiology, 99, 353 368.<br />

Collinge, D. B., Gregersen, P. L. & Thordal-Christensen,<br />

H. (2002). The nature and role of defense response<br />

genes in cereals. In The Powdery Mildews. A<br />

Comprehensive Treatise, ed. R. R. Bélanger, W. R.<br />

Bushnell, A. J. Dik & T. L. W. Carver. St. Paul, USA: APS<br />

Press, pp. 146 160.<br />

Collins, N. C., Sadanandom, A. & Schulze-Lefert, P.<br />

(2002). Genes and molecular mechanisms controlling<br />

powdery mildew resistance in barley. In The<br />

Powdery Mildews. A Comprehensive Treatise, ed. R. R.<br />

Bélanger, W. R. Bushnell, A. J. Dik & T. L. W. Carver. St.<br />

Paul, USA: APS Press, pp. 134 145.<br />

Colot, H. V., Loros, J. J. & Dunlap, J. C. (2005).<br />

Temperature-modulated alternative splicing and<br />

promoter use in the circadian clock gene frequency.<br />

Molecular Biology of the Cell, 16, 5563 5571.<br />

Comménil, P., Belingheri, L. & Dehorter, B. (1998).<br />

Antilipase antibodies prevent infection of <strong>to</strong>ma<strong>to</strong><br />

leaves by Botrytis cinerea. Physiological and Molecular<br />

Plant Pathology, 52, 1 14.<br />

Connick, W. J. & French, R. C. (1991). Volatiles emitted<br />

during the sexual stage of the Canada thistle rust<br />

fungus and by thistle flowers. Journal of Agricultural<br />

and Food Chemistry, 39, 185 188.<br />

Cook, R. T. A., Inman, A. J. & Billings, C. (1997). Identification<br />

and classification of powdery mildew anamorphs<br />

using light and scanning electron microscopy<br />

and host data. Mycological Research, 101, 975 1002.<br />

Cooke, D. E. L., Drenth, A., Duncan, J. M., Wagels, G. &<br />

Brasier, C. M. (2000). A molecular phylogeny of<br />

Phy<strong>to</strong>phthora and related oomycetes. Fungal Genetics<br />

and Biology, 30, 17 32.


REFERENCES<br />

723<br />

Cooke, L. R. & Little, G. (2002). The effect of foliar<br />

application of phosphonate formulations on the<br />

susceptibility of pota<strong>to</strong> tubers <strong>to</strong> late blight. Pest<br />

Management Science, 58, 17 25.<br />

Cooke, R. C. (1964). Ecological characteristics of nema<strong>to</strong>de-trapping<br />

hyphomycetes. II. Germination of<br />

conidia in soil. Annals of Applied Biology, 54, 375 379.<br />

Cooke, R. C. & Godfrey, B. E. S. (1964). A key <strong>to</strong> the<br />

nema<strong>to</strong>de-destroying fungi. Transactions of the British<br />

Mycological Society, 47, 61 74.<br />

Cooke, R. C. & Mitchell, D. T. (1970). Carbohydrate<br />

physiology of sclerotia of Claviceps purpurea during<br />

dormancy and germination. Transactions of the British<br />

Mycological Society, 54, 93 99.<br />

Cooke, R. C. & Rayner, A. D. M. (1984). Ecology of<br />

Saprotrophic <strong>Fungi</strong>. London: Longman.<br />

Cooney, D. G. & Emerson, R. (1964). Thermophilic <strong>Fungi</strong>,<br />

an Account of their Biology, Activities and Classification.<br />

San Francisco: Freeman.<br />

Cooney, E. W., Barr, D. J. S. & Bars<strong>to</strong>w, W. E. (1985). The<br />

ultrastructure of the zoospore of Hyphochytrium<br />

catenoides. Canadian Journal of Botany, 63, 497 505.<br />

Coppin, E., Debuchy, R., Arnaise, S. & Pickard, M. (1997).<br />

Mating types and sexual development in filamen<strong>to</strong>us<br />

ascomycetes. Microbiology and Molecular Biology<br />

Reviews, 61, 411 428.<br />

Coremans-Pelseneer, J. (1973). Isolation of Basidiobolus<br />

meris<strong>to</strong>spores from natural sources. Mycopathologia<br />

et Mycologia Applicata, 49, 173 176.<br />

Corlett, M. (1966). Developmental studies in the<br />

Microascaceae. Canadian Journal of Botany, 44, 79 88.<br />

Corlett, M. (1973). Observations and comments on the<br />

Pleospora centrum type. Nova Hedwigia, 24, 347 360.<br />

Corlett, M. (1991). An annotated list of the published<br />

names in Mycosphaerella and Sphaerella. Mycologia<br />

Memoir, 18, 1 328.<br />

Cormack, B. P., Ghori, N. & Falkow, S. (1999). An<br />

adhesin of the yeast pathogen Candida glabrata<br />

mediating adherence <strong>to</strong> human epithelial cells.<br />

Science, 285, 578 582.<br />

Cornelius, G. & Nakashima, H. (1987). Vacuoles play a<br />

decisive role in calcium homeostasis in Neurospora<br />

crassa. Journal of General Microbiology, 133, 2341 2347.<br />

Corner, E. J. H. (1929). Studies in the morphology of<br />

discomycetes. II. The structure and development of<br />

the ascocarp. Transactions of the British Mycological<br />

Society, 14, 275 291.<br />

Corner, E. J. H. (1932a). The fruit body of Polystictus<br />

xanthopus. Annals of Botany, 46, 71 111.<br />

Corner, E. J. H. (1932b). A Fomes with two systems of<br />

hyphae. Transactions of the British Mycological Society,<br />

17, 51 81.<br />

Corner, E. J. H. (1948). Studies in the basidium. I. The<br />

ampoule effect, with a note on nomenclature. New<br />

Phy<strong>to</strong>logist, 47, 22 51.<br />

Corner, E. J. H. (1950). A Monograph of Clavaria and Allied<br />

Genera. London: Oxford University Press.<br />

Corner, E. J. H. (1953). The construction of polypores. 1.<br />

<strong>Introduction</strong> <strong>to</strong> Polyporus sulphureus, P. squamosus, P.<br />

betulinus and Polystictus microcyclus. Phy<strong>to</strong>morphology, 3,<br />

152 169.<br />

Corrêa, A., Staples, R. C. & Hoch, H. C. (1996). Inhibition<br />

of thigmostimulated cell differentiation with RGDpeptides<br />

in Uromyces germlings. Pro<strong>to</strong>plasma, 194,<br />

91 102.<br />

Cortés, J. C. G., Ishiguro, J., Durán, A. & Ribas, C. (2002).<br />

Localization of the (1,3)b-D-glucan synthase catalytic<br />

subunit homologue Bgs1p/Cps1p from fission yeast<br />

suggests that it is involved in septation, polarized<br />

growth, mating, spore wall formation and spore<br />

germination. Journal of Cell Science, 115, 4081 4096.<br />

Cortesi, P., Gadoury, D. M., Seem, R. C. & Pearson, R. C.<br />

(1995). Distribution and retention of cleis<strong>to</strong>thecia of<br />

Uncinula neca<strong>to</strong>r on the bark of grapevines. Plant<br />

Disease, 79, 15 19.<br />

Cortesi, P., McCulloch, C. E., Song, H., Lin, H. &<br />

Milgroom, M. G. (2001). Genetic control of horizontal<br />

virus transmission in the chestnut blight fungus,<br />

Cryphonectria parasitica. Genetics, 159, 107 118.<br />

Costantin, J. (1936). La culture de la morille et sa forme<br />

conidienne. Annales des Sciences Naturelles Botaniques.<br />

Série 10, 18, 111 129.<br />

Cotter, D. A., Sands, T. W., Virdy, K. J., et al. (1992).<br />

Patterning of development in Dictyostelium discoideum:<br />

fac<strong>to</strong>rs regulating growth, differentiation,<br />

spore dormancy, and germination. Biochemistry and<br />

Cell Biology, 70, 892 919.<br />

Cotty, P. J., Bayman, P., Egel, D. S. & Elias, K. S. (1994).<br />

Agriculture, afla<strong>to</strong>xins and Aspergillus. InThe Genus<br />

Aspergillus. From Taxonomy and Genetics <strong>to</strong> Industrial<br />

Application, ed. K. A. Powell, A. Renwick & J. F.<br />

Peberdy. New York: Plenum Press, pp. 1 27.<br />

Couch, B. C., Fudal, I., Lebrun, M.-H., et al. (2005).<br />

Origins of host-specific populations of the blast<br />

pathogen Magnaporthe oryzae in crop domestication<br />

with subsequent expansion of pandemic clones on<br />

rice and weeds of rice. Genetics, 170, 613 630.<br />

Couch, J. N. (1939). Heterothallism in the Chytridiales.<br />

Journal of the Elisha Mitchell Scientific Society, 55,<br />

409 414.<br />

Couch, J. N. (1945). Observations on the fungus<br />

Catenaria. Mycologia, 37, 163 193.<br />

Couch, J. N. & Bland, C. E., eds. (1985). The Genus<br />

Coelomomyces. Orlando: Academic Press.


724 REFERENCES<br />

Cous<strong>to</strong>u, V., Deleu, C., Saupe, S. & Begueret, J. (1997).<br />

The protein product of the het-s heterokaryon<br />

incompatibility gene of the fungus Podospora<br />

anserina behaves as a prion analog. Proceedings of<br />

the National Academy of Sciences of the USA, 94,<br />

9773 9778.<br />

Cous<strong>to</strong>u-Linares, V., Maddelein, M.-L.,<br />

Bégueret, J. & Saupe, S. (2001). In vivo<br />

aggregation of the HET-s prion protein of the<br />

fungus Podospora anserina. Molecular Microbiology,<br />

42, 1325 1335.<br />

Coutinho, T. A., Rijkenberg, F. H. J. & van Asch, M. A. J.<br />

(1993). Appressorium formation by Hemileia vastatrix.<br />

Mycological Research, 97, 951 956.<br />

Cowen, L. E., Anderson, J. B. & Kohn, L. M. (2002).<br />

Evolution of drug resistance in Candida albicans.<br />

Annual Review of Microbiology, 56, 139 165.<br />

Cox, A. E. & Large, E. C. (1960). Pota<strong>to</strong> blight epidemics<br />

throughout the world. U. S. Department of Agriculture<br />

Handbook, 174. Washing<strong>to</strong>n: USDA.<br />

Cox, R. A. & Magee, D. M. (2004). Coccidioidomycosis:<br />

host response and vaccine development. Clinical<br />

Microbiology Reviews, 17, 804 839.<br />

Craft, C. M. & Nelson, E. B. (1996). Microbial properties<br />

of composts that suppress damping-off and root rot<br />

of creeping bentgrass caused by Pythium graminicola.<br />

Applied and Environmental Microbiology, 62,<br />

1550 1557.<br />

Craig, G. D., Gull, K. & Wood, D. A. (1977). Stipe<br />

elongation in Agaricus bisporus. Journal of General<br />

Microbiology, 102, 337 347.<br />

Craig, G. D., Newsam, R. J., Gull, K. & Wood, D. A. (1979).<br />

An ultrastructural and au<strong>to</strong>radiographic study of<br />

stipe elongation in Agaricus bisporus. Pro<strong>to</strong>plasma, 98,<br />

15 29.<br />

Craigie, J. H. (1927). Discovery of the function<br />

of the pycnia of the rust fungi. Nature, 120,<br />

765 767.<br />

Craigie, J. H. & Green, G. J. (1962). Nuclear behaviour<br />

leading <strong>to</strong> conjugate association in haploid infections<br />

of Puccinia graminis. Canadian Journal of Botany,<br />

40, 163 178.<br />

Cramer, C. L. & Davis, R. H. (1984).<br />

Polyphosphate cation interaction in the amino<br />

acid-containing vacuole of Neurospora crassa. Journal<br />

of Biological Chemistry, 259, 5152 5157.<br />

Creppy, E. E. (2002). Update of survey, regulation and<br />

<strong>to</strong>xic effects of myco<strong>to</strong>xins in Europe. Toxicology<br />

Letters, 127, 19 28.<br />

Crespo Erchiga, V. & Delgado Florencio, V. (2002).<br />

Malassezia species in skin diseases. Current Opinion in<br />

Infectious Diseases, 15, 133 142.<br />

Crittenden, P. D., David, J. C., Hawksworth, D. L. &<br />

Campbell, F. S. (1995). Attempted isolation and<br />

success in the culturing of a broad spectrum of<br />

lichen-forming and lichenicolous fungi. New<br />

Phy<strong>to</strong>logist, 130, 267 297.<br />

Crous, P. W., Aptroot, A., Kang, J. C., Braun, U. &<br />

Wingfield, M. J. (2000). The genus Mycosphaerella and<br />

its anamorphs. Studies in Mycology, 45, 107 121.<br />

Crous, P. W., Kang, J.-C. & Braun, U. (2001). A phylogenetic<br />

redefinition of anamorph genera in<br />

Mycosphaerella based on ITS rDNA sequence and<br />

morphology. Mycologia, 93, 1081 1101.<br />

Crowe, F. J., Hall, D. H., Greathead, A. S. & Baghott, K. G.<br />

(1980). Inoculum density of Sclerotium cepivorum and<br />

the incidence of white rot of onion and garlic.<br />

Phy<strong>to</strong>pathology, 70, 64 69.<br />

Croxall, H. E., Collingwood, C. A. & Jenkins, J. E. E.<br />

(1951). Observations on brown rot (Sclerotinia fructigena)<br />

of apple in relation <strong>to</strong> injury by earwigs<br />

(Forficula auricularia). Annals of Applied Biology, 38,<br />

833 843.<br />

Crute, I. R. (1984). The integrated use of genetic and<br />

chemical methods for the control of lettuce downy<br />

mildew (Bremia lactucae). Crop Protection, 3, 223 242.<br />

Crute, I. R. & Dixon, G. R. (1981). Downy mildew<br />

diseases caused by the genus Bremia. InThe Downy<br />

Mildews, ed. D. M. Spencer. London: Academic Press,<br />

pp. 421 460.<br />

Cruz, M. C., del Poeta, M., Wang, P., et al. (2000).<br />

Immunosuppressive and nonimmunosuppressive<br />

cyclosporine analogs are <strong>to</strong>xic <strong>to</strong> the opportunistic<br />

fungal pathogen Cryp<strong>to</strong>coccus neoformans via cyclophilin-dependent<br />

inhibition of calcineurin.<br />

Antimicrobial Agents and Chemotherapy, 44, 143 149.<br />

Cuda, J. P., Hornby, J. A., Cotterill, B. & Cattell, M. (1997).<br />

Evaluation of Lagenidium giganteum for biological<br />

control of Mansonia mosqui<strong>to</strong>es in Florida (Diptera:<br />

Culicidae). Biological Control, 8, 124 130.<br />

Cuenca-Estrella, M., Gomez-Lopez, A., Mellado, E., et al.<br />

(2003). Scopulariopsis brevicaulis, a fungal pathogen<br />

resistant <strong>to</strong> broad-spectrum antifungal agents.<br />

Antimicrobial Agents and Chemotherapy, 47, 2339 2341.<br />

Cui, J. & Chisti, Y. (2003). Polysaccharopeptides of<br />

Coriolus versicolor: physiological activity, uses, and<br />

production. Biotechnology Advances, 21, 109 122.<br />

Culberson, C. F. & Ahmadjian, V. (1980). Artificial<br />

reestablishment of lichens. II: Secondary products of<br />

resynthesised Cladonia cristatella and Lecanora chrysoleuca.<br />

Mycologia, 72, 90 109.<br />

Cullum, F. J. & Webster, J. (1977). Cleis<strong>to</strong>carp dehiscence<br />

in Phyllactinia. Transactions of the British<br />

Mycological Society, 68, 316 320.


REFERENCES<br />

725<br />

Culvenor, C. C., Beck, A. B., Clarke, M., et al. (1977).<br />

Isolation of <strong>to</strong>xic metabolites of Phomopsis lep<strong>to</strong>stromiformis<br />

responsible for lupinosis. Australian Journal<br />

of Biological Science, 30, 269 277.<br />

Cummins, G. B. & Hiratsuka, Y. (2003). Illustrated Genera<br />

of Rust <strong>Fungi</strong>, 3rd edn. St. Paul, USA: APS Press.<br />

Cunfer, B. M. (2000). Stagonospora and Sep<strong>to</strong>ria diseases<br />

of barley, oat, and rye. Canadian Journal of Plant<br />

Pathology, 22, 332 348.<br />

Cunfer, B. M. & Seckinger, A. (1977). Survival of Claviceps<br />

purpurea and C. paspali sclerotia. Mycologia, 69,<br />

1142 1148.<br />

Cunfer, B. M. & Ueng, P. P. (1999). Taxonomy and<br />

identification of Sep<strong>to</strong>ria and Stagonospora species on<br />

small-grain cereals. Annual Review of Phy<strong>to</strong>pathology,<br />

37, 267 284.<br />

Currah, R. S. (1985). Taxonomy of the Onygenales:<br />

Arthrodermataceae, Gymnoascaceae,<br />

Myxotrichaceae and Onygenaceae. Mycotaxon, 24,<br />

1 216.<br />

Currah, R. S. (1994). Peridial morphology and evolution<br />

in the pro<strong>to</strong>tunicate ascomycetes. In Ascomycete<br />

Systematics. Problems and Perspectives in the Nineties, ed.<br />

D. L. Hawksworth. New York: Academic Press,<br />

pp. 281 293.<br />

Currah, R. S., Smreciu, E. A., Lehesvirta, T., Niemi, M. &<br />

Larsen, K. W. (2000). <strong>Fungi</strong> in the winter diets of<br />

northern flying squirrels and red squirrels in the<br />

boreal mixedwood forest of northeastern Alberta.<br />

Canadian Journal of Botany, 78, 1514 1520.<br />

Curtis, F. C., Evans, G. H., Lillis, V., Lewis, D. H. & Cooke,<br />

R. C. (1978). Studies on Mucoralean parasites. I. Some<br />

effects of Pip<strong>to</strong>cephalis species on host growth. New<br />

Phy<strong>to</strong>logist, 80, 157 165.<br />

Curtis, K. M. (1921). The life his<strong>to</strong>ry and cy<strong>to</strong>logy of<br />

Synchytrium endobioticum (Schilb.) Perc., the cause of<br />

wart disease in pota<strong>to</strong>. Philosophical Transactions of the<br />

Royal Society, B 210, 409 478.<br />

Curtis, M. J. & Wolpert, T. J. (2002). The oat mi<strong>to</strong>chondrial<br />

permeability transition and its implication in<br />

vic<strong>to</strong>rin binding and induced cell death. Plant<br />

Journal, 29, 295 312.<br />

Cushion, M. T. (2004). Pneumocystis: unraveling the<br />

cloak of obscurity. Trends in Microbiology, 12,<br />

243 249.<br />

Czeczuga, B. (1983). Muta<strong>to</strong>xanthin, the dominant<br />

carotenoid in lichens of the Xanthoria genus.<br />

Biochemical Systematics and Ecology, 11, 329 331.<br />

Czymmek, K. J., Bourett, T. M. & Howard, R. J. (1996).<br />

Immunolocalization of tubulin and actin in thicksectioned<br />

fungal hyphae and methacrylate deembedment.<br />

Journal of Microscopy, 181, 153 161.<br />

Daboussi, M.-J. & Capy, P. (2003). Transposable elements<br />

in filamen<strong>to</strong>us fungi. Annual Review of Microbiology,<br />

57, 275 299.<br />

Dahlberg, A. (1997). Population ecology of Suillus<br />

variegatus in old Swedish Scots pine forests.<br />

Mycological Research, 101, 47 54.<br />

Dahlberg, A. & Finlay, R. D. (1999). Suillus. In<br />

Ec<strong>to</strong>mycorrhizal <strong>Fungi</strong>. Key Genera in Profile, ed. J. W. G.<br />

Cairney & S. M. Chambers. Berlin: Springer-Verlag,<br />

pp. 33 64.<br />

Dahlberg, A. & Stenlid, J. (1990). Population structure<br />

and dynamics in Suillus bovinus as indicated by<br />

spatial distribution of fungal clones. New Phy<strong>to</strong>logist,<br />

115, 487 493.<br />

Dahlberg, A. & Stenlid, J. (1994). Size, distribution and<br />

biomass of genets in populations of Suillus bovinus (L.:<br />

Fr.) Roussel revealed by somatic incompatibility. New<br />

Phy<strong>to</strong>logist, 128, 225 234.<br />

Dahlstrom, J. L., Smith, J. E. & Weber, N. S. (2000).<br />

Mycorrhiza-like interaction by Morchella species with<br />

species of the Pinaceae in pure culture synthesis.<br />

Mycorrhiza, 9, 279 285.<br />

Daly, R. & Hearn, M. T. W. (2005). Expression of heterologous<br />

proteins in Pichia pas<strong>to</strong>ris: a useful experimental<br />

<strong>to</strong>ol in protein engineering and production.<br />

Journal of Molecular Recognition, 18, 119 138.<br />

Danell, E. (1994). Formation and growth of the ec<strong>to</strong>mycorrhiza<br />

of Cantharellus cibarius. Mycorrhiza, 5, 89 97.<br />

Daniels, A., Lucas, J. A. & Peberdy, J. F. (1991).<br />

Morphology and ultrastructure of W and R pathotypes<br />

of Pseudocercosporella herpotrichoides on wheat<br />

seedlings. Mycological Research, 95, 385 397.<br />

Daniels, A., Papaikonomou, M., Dyer, P. S. & Lucas, J. A.<br />

(1995). Infection of wheat seedlings by ascospores of<br />

Tapesia yallundae: morphology of the infection<br />

process and evidence for recombination.<br />

Phy<strong>to</strong>pathology, 85, 918 927.<br />

Daniels, J. A. (1961). Chae<strong>to</strong>mium piluliferum sp. nov., the<br />

perfect state of Botryotrichum piluliferum. Transactions<br />

of the British Mycological Society, 44, 79 86.<br />

Dao, D. N., Kessin, R. H. & Ennis, H. L. (2000).<br />

Developmental cheating and the evolutionary biology<br />

of Dictyostelium and Myxococcus. Microbiology, 146,<br />

1505 1512.<br />

Datema, R., van den Ende, H. & Wessels, J. G. H. (1977).<br />

The hyphal wall of Mucor mucedo. 1. Polyanionic<br />

polymers. European Journal of Biochemistry, 80,<br />

611 619.<br />

Daub, M. E. & Ehrenshaft, M. (2000). The pho<strong>to</strong>activated<br />

Cercospora <strong>to</strong>xin cercosporin: contributions <strong>to</strong> plant<br />

disease and fundamental biology. Annual Review of<br />

Phy<strong>to</strong>pathology, 38, 461 490.


726 REFERENCES<br />

Davey, C. B. & Papavizas, G. C. (1962). Growth and<br />

sexual reproduction of Aphanomyces euteiches as<br />

affected by the oxidation state of sulfur. American<br />

Journal of Botany, 49, 400 404.<br />

Davey, J. (1992). Mating pheromones of the fission yeast<br />

Schizosaccharomyces pombe: purification and structural<br />

characterization of M-fac<strong>to</strong>r and isolation and<br />

analysis of two genes encoding the pheromone.<br />

EMBO Journal, 11, 951 960.<br />

Davey, J. (1998). Fusion of a fission yeast. Yeast, 14,<br />

1529 1566.<br />

David, M., Gabriel, M. & Kopecká, M. (2003). Unusual<br />

ultrastructural characteristics of the yeast Malassezia<br />

pachydermatis. Scripta Medica, 76, 173 186.<br />

Davidse, L. C. & Ishii, H. (1995). Biochemical and<br />

molecular aspects of the mechanisms of action of<br />

benzimidazoles, N-phenylcarbamates and N-phenylformamidoximes<br />

and the mechanisms of resistance<br />

<strong>to</strong> these compounds in fungi. In Modern Selective<br />

<strong>Fungi</strong>cides. Properties, Applications, Mechanisms of Action,<br />

2nd edn, ed. H. Lyr. New York: Gustav Fischer Verlag,<br />

pp. 305 322.<br />

Davidse, L. C., Hofman, A. E. & Velthuis, G. C. M. (1983).<br />

Specific interference of metalaxyl with endogenous<br />

RNA polymerase activity in isolated nuclei from<br />

Phy<strong>to</strong>phthora megasperma f. sp. medicaginis.<br />

Experimental Mycology, 7, 344 361.<br />

Davidse, L. C., van den Berg-Velthuis, G. C. M., Mantel,<br />

B. C. & Jespers, A. B. K. (1991). Phenylamides and<br />

Phy<strong>to</strong>phthora. InPhy<strong>to</strong>phthora, ed. J. A. Lucas, R. C.<br />

Shat<strong>to</strong>ck, D. S. Shaw & L. R. Cooke. Cambridge:<br />

Cambridge University Press, pp. 349 360.<br />

Davis, E. E. (1967). Zygospore formation in Syzygites<br />

megalocarpus. Canadian Journal of Botany, 45, 531 532.<br />

Davis, R. H. (1995). The genetics of Neurospora. In The<br />

Mycota II: Genetics and Biotechnology, ed. W. Kück.<br />

Berlin: Springer-Verlag, pp. 3 18.<br />

Davoli, P., Mierau, V. & Weber, R. W. S. (2004).<br />

Carotenoids and fatty acids in red yeasts<br />

Sporobolomyces roseus and Rhodo<strong>to</strong>rula glutinis. Applied<br />

Biochemistry and Microbiology, 40, 460 465.<br />

Dawe, A. L. & Nuss, D. L. (2001). Hypoviruses and<br />

chestnut blight: exploiting viruses <strong>to</strong> understand<br />

and modulate fungal pathogenesis. Annual Review of<br />

Genetics, 35, 1 29.<br />

Day, A. R., Poon, N. H. & Stewart, G. G. (1975). Fungal<br />

fimbriae. III. The effect on flocculation in<br />

Saccharomyces. Canadian Journal of Microbiology, 21,<br />

558 564.<br />

Day, A. W. (1998). Nonmeiotic mechanisms of recombination<br />

in the anther smut Microbotryum violaceum.<br />

International Journal of Plant Sciences, 159, 185 191.<br />

Day, A. W. & Garber, E. D. (1988). Ustilago violacea, anther<br />

smut of the Caryophyllaceae. Advances in Plant<br />

Pathology, 6, 457 482.<br />

Day, P. R. (1974). Genetics of Host Parasite Interaction. San<br />

Francisco: W. H. Freeman & Co.<br />

Day, P. R. (1981). Fungal virus populations in corn smut<br />

from Connecticut. Mycologia, 73, 379 391.<br />

Deacon, J. W. & Donaldson, S. P. (1993).<br />

Molecular recognition in the homing responses of<br />

zoosporic fungi, with special reference <strong>to</strong><br />

Pythium and Phy<strong>to</strong>phthora. Mycological Research,<br />

97, 1153 1171.<br />

Deacon, J. W., Laing, S. A. K. & Berry, L. A. (1990). Pythium<br />

mycoparasiticum sp. nov., an aggressive mycoparasite<br />

from British soils. Mycotaxon, 42, 1 8.<br />

Dean, R. A. (1997). Signal pathways and appressorium<br />

morphogenesis. Annual Review of Phy<strong>to</strong>pathology, 35,<br />

211 234.<br />

de Baets, S. & Vandamme, E. J. (2001). Extracellular<br />

Tremella polysaccharides: structure, properties and<br />

applications. Biotechnology Letters, 23, 1361 1366.<br />

de Bary, H. A. (1853) Untersuchungen über die Brandpilze<br />

und die durch sie verursachten Krankheiten der Pflanzen<br />

mit Rücksicht auf das Getreide und andere Nutzpflanzen.<br />

Berlin: Müller.<br />

de Bary, A. (1866). Morphologie und Physiologie der Pilze,<br />

Flechten und Myxomyceten. Leipzig: Wilhelm<br />

Engelmann.<br />

de Bary, A. (1887). Comparative Morphology and Biology of<br />

the <strong>Fungi</strong>, Myce<strong>to</strong>zoa and Bacteria, 2nd edn (English<br />

translation). Oxford: Clarendon Press.<br />

Debuchy, R. & Coppin, E. (1992). The mating types of<br />

Podospora anserina: functional analysis and sequence<br />

of the fertilization domains. Molecular and General<br />

Genetics, 233, 113 121.<br />

Deckert, R. J., Melville, L. H. & Peterson, R. L. (2001).<br />

Structural features of a Lophodermium endophyte<br />

during the cryptic life-cycle phase in the<br />

foliage of Pinus strobus. Mycological Research,<br />

105, 991 997.<br />

de Cock, A. W. A. M., Mendoza, L., Padhye, A. A., Ajello, L.<br />

& Prell, H. H. (1987). Pythium insidiosum sp. nov., the<br />

etiologic agent of pythiosis. Journal of Clinical<br />

Microbiology, 25, 344 349.<br />

Decottignies, A., Sanchez-Perez, I. & Nurse, P. (2003).<br />

Schizosaccharomyces pombe essential genes: a pilot<br />

study. Genome Research, 13, 399 406.<br />

Degawa, Y. & Tokamasu, S. (1997). Zygospore formation<br />

in Mortierella capitata. Mycoscience, 38, 384 394.<br />

Degawa, Y. & Tokamasu, S. (1998). Zygospore formation<br />

in Mortierella umbellata. Mycological Research, 102,<br />

593 598.


REFERENCES<br />

727<br />

Degousée, N., Gupta, G. D., Lew, R. R. & Heath, I. B.<br />

(2000). A putative spectrin-containing membrane<br />

skele<strong>to</strong>n in hyphal tips of Neurospora crassa. Fungal<br />

Genetics and Biology, 30, 33 44.<br />

De Groot, P. W. J., Visser, J., Van Griensven, L. J. L. D. &<br />

Schaap, P. J. (1998). Biochemical and molecular<br />

aspects of growth and fruiting of the edible mushroom<br />

Agaricus bisporus. Mycological Research, 102,<br />

1297 1308.<br />

de Hoog, G. S. (1974). The genera Blas<strong>to</strong>botrys, Sporothrix,<br />

Calcarisporium and Calcarisporiella gen. nov. Studies in<br />

Mycology, 7, 1 84.<br />

de Hoog, G. S. (1977). Rhinocladiella and allied genera.<br />

Studies in Mycology, 15, 1 140.<br />

de Hoog, G. S. & Scheffer, R. J. (1984). Cera<strong>to</strong>cystis versus<br />

Ophios<strong>to</strong>ma: a reappraisal. Mycologia, 76, 292 299.<br />

de Hoog, G. S. & Smith, M. T. (2004). Ribosomal gene<br />

phylogeny and species delimitation in Geotrichum<br />

and its teleomorphs. Studies in Mycology, 50, 489 515.<br />

de Hoog, G. S., Kurtzman, C. P., Phaff, H. J. & Miller,<br />

M. W. (1998). Eremothecium Borzi emend. Kurtzman.<br />

In The Yeasts: A Taxonomic Study, 4th edn, ed. C. P.<br />

Kurtzman & J. W. Fell. Elsevier: Amsterdam,<br />

pp. 201 208.<br />

de Hoog, G. S., Guarro, J., Gené, J. & Figueras, M. J.<br />

(2000a). Atlas of Clinical <strong>Fungi</strong>, 2nd edn. Baarn,<br />

Netherlands: Centraalbureau voor<br />

Schimmelcultures; Reus, Spain: Universitat Rovira i<br />

Virgili.<br />

de Hoog, G. S., Queiroz-Telles, F., Haase, G., et al. (2000b).<br />

Black fungi: clinical and pathogenic approaches.<br />

Medical Mycology, 38, Supplement 1, 243 250.<br />

Deising, H. B., Nicholson, R. B., Haug, M., Howard, R. J. &<br />

Mendgen, K. (1992). Adhesion pad formation and the<br />

involvement of cutinase and esterases in the<br />

attachment of uredospores <strong>to</strong> the host cuticle. Plant<br />

Cell, 4, 1101 1111.<br />

de Jong, E., Field, J. A. & de Bont, J. A. M. (1994a). Aryl<br />

alcohols in the physiology of ligninolytic fungi. FEMS<br />

Microbiology Reviews, 13, 153 187.<br />

de Jong, E., Field, J. A., Spinnler, H. E., Wijnberg,<br />

J. B. P. A. & de Bont, J. A. M. (1994b). Significant<br />

biogenesis of chlorinated aromatics by fungi in<br />

natural environments. Applied and Environmental<br />

Microbiology, 60, 264 270.<br />

de Jong, J. C., McCormack, B. J., Smirnoff, N. & Talbot,<br />

N. J. (1997). Glycerol generates turgor in rice blast.<br />

Nature, 389, 244 245.<br />

Dekhuijzen, H. M. (1980). The occurrence of free and<br />

bound cy<strong>to</strong>kinins in clubroots and Plasmodiophora<br />

brassicae infected turnip tissue cultures. Physiologia<br />

Plantarum, 49, 169 176.<br />

de la Torre, M. J., Millan, M. C., Perez-Juan, P., Morales, J.<br />

& Ortega, J. M. (1999). Indigenous yeasts associated<br />

with two Vitis vinifera grape varieties cultured in<br />

southern Spain. Microbios, 100, 27 40.<br />

Dell, B., Malajczuk, N. & Dunstan, W. A. (2002).<br />

Persistence of some Australian Pisolithus species<br />

introduced in<strong>to</strong> eucalypt plantations in China. Forest<br />

Ecology and Management, 169, 271 281.<br />

Delmas, J. (1978). Tuber species. In The Biology and<br />

Cultivation of Edible Mushrooms, ed. S. T. Chang & W. A.<br />

Hayes. New York: Academic Press, pp. 645 681.<br />

Demain, A. L. & Elander, R. P. (1999). The b-lactam<br />

antibiotics: past, present, and future. An<strong>to</strong>nie van<br />

Leeuwenhoek, 75, 5 19.<br />

Dengis, P. B. & Rouxhet, P. G. (1997). Surface properties<br />

of <strong>to</strong>p- and bot<strong>to</strong>m-fermenting yeast. Yeast, 13,<br />

931 943.<br />

Dennis, R. L. (1970). A middle Pennsylvanian basidiomycete<br />

mycelium with clamp connections.<br />

Mycologia, 62, 578 584.<br />

Dennis, R. W. G. (1960). British Cup <strong>Fungi</strong> and Their Allies.<br />

London: Ray Society.<br />

Dennis, R. W. G. (1981). British Ascomycetes. Vaduz: J.<br />

Cramer.<br />

de Nobel, H. & Barnett, J. A. (1991). Passage of molecules<br />

through yeast cell walls: a brief essay-review. Yeast, 7,<br />

313 323.<br />

de Nobel, H., Sietsma, J. H., van den Ende, H. & Klis,<br />

F. M. (2001). Molecular organization and construction<br />

of the fungal cell wall. In The Mycota VIII: Biology<br />

of the Fungal Cell, ed. R. J. Howard & N. A. R. Gow.<br />

Berlin: Springer-Verlag, pp. 181 200.<br />

Descals, E. & Webster, J. (1984). Branched aquatic<br />

conidia in Erynia and En<strong>to</strong>mophthora sensu la<strong>to</strong>.<br />

Transactions of the British Mycological Society, 83,<br />

669 682.<br />

Descals, E., Webster, J., Ladle, M. & Bass, J. A. B. (1981).<br />

Variations in asexual reproduction in species of<br />

En<strong>to</strong>mophthora on aquatic insects. Transactions of the<br />

British Mycological Society, 77, 85 102.<br />

Descals, E., Marvanová, L. & Webster, J. (1998). New taxa<br />

and combinations in aquatic hyphomycetes.<br />

Canadian Journal of Botany, 76, 1647 1659.<br />

Deshpande, M. S., Rale, V. B. & Lynch, J. M. (1992).<br />

Aureobasidium pullulans in applied microbiology: a<br />

status report. Enzyme and Microbial Technology, 14,<br />

514 527.<br />

de Silva, L., Youatt, J., Gooday, G. R. & Gow, N. A. R.<br />

(1992). Inwardly directed ionic currents of Allomyces<br />

macrogynus and other water moulds indicate sites of<br />

pro<strong>to</strong>n-driven nutrient transport but are incidental<br />

<strong>to</strong> tip growth. Mycological Research, 96, 925 931.


728 REFERENCES<br />

de Vrije, T., An<strong>to</strong>ine, N., Buitelaar, R. M., et al. (2001).<br />

The fungal biocontrol agent Coniothyrium minitans:<br />

production by solid-state fermentation, application<br />

and marketing. Applied Microbiology and Biotechnology,<br />

56, 58 68.<br />

De Wolf, E. D., Effertz, R. J., Ali, S. & Francl, L. J. (1998).<br />

Vistas of tan spot research. Canadian Journal of Plant<br />

Pathology, 20, 349 370.<br />

deZwaan, T. M., Carroll, A. M., Valent, B. &<br />

Sweigard, J. A. (1999). Magnaporthe grisea Pth11p<br />

is a novel plasma membrane protein that mediates<br />

appressorium differentiation in response<br />

<strong>to</strong> inductive substrate cues. Plant Cell, 11,<br />

2013 2030.<br />

Dhaliwal, H. S. (1989). Multiplication of secondary<br />

sporidia of Tilletia indica on soil and wheat leaves and<br />

spikes and incidence of Karnal bunt. Canadian Journal<br />

of Botany, 67, 2387 2390.<br />

Dhingra, O. D., Mizubuti, E. S. G. & Santana, F. M. (2003).<br />

Chae<strong>to</strong>mium globosum for reducing primary inoculum<br />

of Diaporthe phaseolorum f. sp. meridionalis in soilsurface<br />

soybean stubble in field conditions. Biological<br />

Control, 26, 302 310.<br />

Dick, M. W. (1970). Saprolegniaceae on insect exuviae.<br />

Transactions of the British Mycological Society, 55,<br />

449 459.<br />

Dick, M. W. (1983). Validation of the class name<br />

Hyphochytriomycetes. In Zoosporic Plant<br />

Pathogens, ed. S. T. Buczacki. London: Academic<br />

Press, p. 285.<br />

Dick, M. W. (1990a). Phylum Oomycota. In Handbook of<br />

Pro<strong>to</strong>ctista, ed. L. Margulis, J. O. Corliss, M. Melkonian<br />

& D. J. Chapman. Bos<strong>to</strong>n: Jones & Bartlett,<br />

pp. 661 685.<br />

Dick, M. W. (1990b). Keys <strong>to</strong> Pythium. Reading, UK:<br />

University of Reading.<br />

Dick, M. W. (1995). Sexual reproduction in the<br />

Peronosporomycetes (chromistan fungi). Canadian<br />

Journal of Botany, 73, S712 S724.<br />

Dick, M. W. (1997). <strong>Fungi</strong>, flagella and phylogeny.<br />

Mycological Research, 101, 385 394.<br />

Dick, M. W. (1998). The species and systematic<br />

position of Crypticola in the Peronosporomycetes,<br />

and new names for the genus Halocrusticidia and<br />

species therein. Mycological Research, 102,<br />

1062 1066.<br />

Dick, M. W. (2001a). Straminipilous <strong>Fungi</strong>. Dordrecht:<br />

Kluwer Academic Publishers.<br />

Dick, M. W. (2001b). The Peronosporomycetes. In The<br />

Mycota VIIA: Systematics and Evolution, ed. D. J.<br />

McLaughlin, E. G. McLaughlin & P. A. Lemke. Berlin:<br />

Springer-Verlag, pp. 39 72.<br />

Dick, M. W. (2002). Towards an understanding of the<br />

evolution of the downy mildews. In Advances in Downy<br />

Mildew Research, ed. P. T. N. Spencer-Phillips, U. Gisi &<br />

A. Labeda. Dordrecht: Kluwer Academic Publishers,<br />

pp. 1 57.<br />

Dick, M. W. (2003). The Peronosporomycetes and other<br />

flagellate fungi. In Pathogenic <strong>Fungi</strong> in Humans and<br />

Animals, 2nd edn, ed. D. H. Howard. New York: Marcel<br />

Dekker, pp. 17 66.<br />

Dick, M. W., Vick, M. C., Gibbings, J. G., Hedderson, T. A.<br />

& Lopez Lastra, C. C. (1999). 18S rDNA for species of<br />

Lep<strong>to</strong>legnia and other Peronosporomycetes: justification<br />

for the subclass taxa Saprolegniomycetidae and<br />

Peronosporomycetidae and division of the<br />

Saproleniaceae sensu la<strong>to</strong> in<strong>to</strong> the families<br />

Lep<strong>to</strong>legniaceae and Saprolegniaceae. Mycological<br />

Research, 103, 1119 1125.<br />

Dickinson, C. H. & Greenhalgh, J. R. (1977). Host range<br />

and taxonomy of Peronospora on crucifers.<br />

Transactions of the British Mycological Society, 69,<br />

111 116.<br />

Diederich, P. (1996). The lichenicolous<br />

Heterobasidiomycetes. Bibliotheca Mycologica, 61,<br />

1 198. Stuttgart: Gebr. Borntraeges.<br />

Diéguez-Uribeondo, J., Gierz, G. & Bartnicki-García, S.<br />

(2004). Image analysis of hyphal morphogenesis in<br />

Saprolegniaceae (Oomycetes). Fungal Genetics and<br />

Biology, 41, 293 307.<br />

Dietel, P. (1928). Reihe Uredinales. In Die Natürlichen<br />

Pflanzenfamilien 6, 2nd edn, ed. A. Engler. Leipzig: W.<br />

Engelmann, pp. 24 98.<br />

Dijksterhuis, J., Veenhuis, M. & Harder, W. (1990).<br />

Ultrastructural study of adhesion and initial stages<br />

of infection of nema<strong>to</strong>des by conidia of Drechmeria<br />

coniospora. Mycological Research, 94, 1 8.<br />

Dijksterhuis, J., Harder, W., Wyss, U. & Veenhuis, M.<br />

(1991). Colonization and digestion of nema<strong>to</strong>des by<br />

the endoparasitic nema<strong>to</strong>phagous fungus Drechmeria<br />

coniospora. Mycological Research, 95, 873 878.<br />

Dijksterhuis, J., Harder, W. & Veenhuis, M. (1993).<br />

Proliferation and function of microbodies in the<br />

nema<strong>to</strong>phagous fungus Arthrobotrys oligospora during<br />

growth on oleic acid or D-alanine as the sole carbon<br />

source. FEMS Microbiology Letters, 112, 125 130.<br />

Dimond, A. E. (1966). Effectiveness of antibiotics<br />

against forest tree rusts: a summary of present<br />

status. Journal of Forestry, 64, 379 382.<br />

Di Pietro, A., Madrid, M. P., Caracuel, Z., Delgado-<br />

Jarana, J. & Roncero, M. I. G. (2003). Fusarium<br />

oxysporum: exploring the molecular arsenal of a<br />

vascular wilt fungus. Molecular Plant Pathology, 4,<br />

315 325.


REFERENCES<br />

729<br />

Dissing, H. (1986). The genus Helvella in Europe with<br />

special emphasis on the species found in Norden.<br />

Dansk Botanisk Arkiv, 25, 1 172.<br />

Dissing, H., Eckblad, F.-E. & Lange, M. (2000).<br />

Pezizales. In Nordic Macromycetes. 1. Ascomycetes, ed.<br />

L. Hansen & H. Knudsen. Copenhagen: Nordswamp,<br />

pp. 55 127.<br />

Dittrich, H. H. (1995). Wine and brandy. In Biotechnology<br />

9: Enzymes, Biomass, Food and Feed, ed. G. Reed & T. W.<br />

Nagodawithana. Weinheim, Germany: VCH,<br />

pp. 463 504.<br />

Dix, N. J. & Webster, J. (1995). Fungal Ecology. London:<br />

Chapman & Hall.<br />

Dixon, D. M. (2001). Coccidioides immitis as a select agent<br />

of bioterrorism. Journal of Applied Microbiology, 91,<br />

602 605.<br />

Dixon, K. P., Xu, J.-R., Smirnoff, N. & Talbot, N. J.<br />

(1999). Independent signaling pathways regulate<br />

cellular turgor during hyperosmotic stress<br />

and appressorium-mediated plant<br />

infection by Magnaporthe grisea. Plant Cell,<br />

11, 2045 2058.<br />

Dodd, J. L. & McCracken, D. A. (1972). Starch in fungi. Its<br />

molecular structure in three genera and an hypothesis<br />

concerning its physiological role. Mycologia, 64,<br />

1341 1343.<br />

Dodge, B. O. (1937). The perithecial cavity<br />

formation in a Lep<strong>to</strong>sphaeria on Opuntia. Mycologia, 29,<br />

707 716.<br />

Doggett, M. S. & Porter, D. (1996). Sexual reproduction<br />

in the fungal parasite, Zygorhizidium plank<strong>to</strong>nicum.<br />

Mycologia, 88, 720 732.<br />

Doguet, G. (1962). Digita<strong>to</strong>spora marina n.g., n.sp.,<br />

basidiomycète marin. Comptes Rendus Hebdomadaires<br />

des Séances de l’Académie des Sciences, D 254,<br />

4336 4338.<br />

Doguet, G. (1967). Nia vibrissa Moore et Myers, remarquable<br />

basidiomycète marin. Comptes Rendus<br />

Hebdomadaires des Séances de l’Académie des Sciences, D<br />

265, 1780 1783.<br />

Doling, D. A. (1966). Loose smut in wheat and barley.<br />

Agriculture, 73, 523 527.<br />

Domnas, A., Jaronski, S. & Han<strong>to</strong>n, W. K. (1986). The<br />

zoospore and flagellar mastigonemes of Lagenidium<br />

giganteum (Oomycetes, Lagenidiales). Mycologia, 78,<br />

810 817.<br />

Domsch, K. H., Gams, W. & Anderson, T.-H. (1980).<br />

Compendium of Soil <strong>Fungi</strong>. London: Academic Press.<br />

Donaldson, S. P. & Deacon, J. W. (1993). Changes in<br />

motility of Pythium zoospores induced by calcium<br />

and calcium-modulating drugs. Mycological Research,<br />

97, 877 883.<br />

Dong, J. O., Chen, W. & Crane, J. L. (1998). Phylogenetic<br />

studies of the Lep<strong>to</strong>sphaeriaceae, Pleosporaceae and<br />

some other Loculoascomycetes based on nuclear<br />

ribosomal DNA sequences. Mycological Research, 102,<br />

151 156.<br />

Donk, M. A. (1964). A conspectus of the families of the<br />

Aphyllophorales. Persoonia, 3, 199 324.<br />

Donofrio, N. M. & Delaney, T. P. (2001). Abnormal<br />

callose response phenotype and hypersusceptibility<br />

<strong>to</strong> Peronospora parasitica in defence-compromised<br />

Arabidopsis nim1 1 and salicylate hydroxylaseexpressing<br />

plants. Molecular Plant Microbe<br />

Interactions, 14, 439 450.<br />

Doss, R. P. (1999). Composition and enzymatic activity<br />

of the extracellular matrix secreted by germlings of<br />

Botrytis cinerea. Applied and Environmental Microbiology,<br />

65, 404 408.<br />

Doss, R. P., Potter, S. W., Soeldner, A. H., Christian, J. K.<br />

& Fukunaga, L. E. (1995). Adhesion of germlings of<br />

Botrytis cinerea. Applied and Environmental Microbiology,<br />

61, 260 265.<br />

Doss, R. P., Potter, S. W., Christian, J. K.,<br />

Soeldner, A. H. & Chastagner, G. A. (1997). The<br />

conidial surface of Botrytis cinerea and several<br />

other Botrytis species. Canadian Journal of Botany,<br />

75, 612 617.<br />

Doster, M. A. & Fry, W. E. (1991). Evaluation by<br />

computer simulation of strategies <strong>to</strong> time metalaxyl<br />

applications for improved control of pota<strong>to</strong> late<br />

blight. Crop Protection, 10, 209 214.<br />

Dowding, E. S. & Bulmer, G. S. (1964). Notes on the<br />

cy<strong>to</strong>logy and sexuality of puffballs. Canadian Journal<br />

of Microbiology, 10, 783 789.<br />

Dowding, P. & Richardson, D. H. S. (1990). Leafyeasts as<br />

indica<strong>to</strong>rs of air quality in Europe. Environmental<br />

Pollution, 66, 223 235.<br />

Dowe, A. (1987). Räuberische Pilze. Wittenberg: A.<br />

Ziemsen Verlag.<br />

Dowley, L. J., Bannon, E., Cooke, L. R., Keane, T. &<br />

O’Sullivan, E., eds. (1995). Phy<strong>to</strong>phthora infestans 150.<br />

Dublin: Boole Press.<br />

Dowsett, J. A., Reid, J. & van Caeseele, L. (1977).<br />

Transmission and scanning electron microscope<br />

observations on the trapping of nema<strong>to</strong>des by<br />

Dactylaria brochopaga. Canadian Journal of Botany, 55,<br />

2945 2955.<br />

Dray<strong>to</strong>n, F. L. & Groves, J. W. (1952). Stromatinia narcissi,<br />

a new sexually dimorphic discomycete. Mycologia, 48,<br />

655 676.<br />

Drechsler, C. (1956). Supplementary developmental<br />

stages of Basidiobolus ranarum and Basidiobolus<br />

hap<strong>to</strong>sporus. Mycologia, 48, 655 676.


730 REFERENCES<br />

Drechsler, C. (1960). Two root rot fungi closely related<br />

<strong>to</strong> Pythium ultimum. Sydowia, 14, 107 115.<br />

Drechsler, C. (1962). A nema<strong>to</strong>de-capturing phycomycete<br />

with distally adhesive branches and proximally<br />

imbedded fusiform conidia. American Journal of<br />

Botany, 49, 1089 1095.<br />

Dreyfuss, M., Härri, E., Hofmann, H., et al. (1976).<br />

Cyclosporin A and C; new metabolites from<br />

Trichoderma polysporum (Link ex Pers.) Rifai. European<br />

Journal of Applied Microbiology, 3, 125 133.<br />

Dring, D. M. (1973). Gasteromycetes. In The <strong>Fungi</strong>, an<br />

Advanced Treatise IVB: A Taxonomic Review with Keys, ed.<br />

G. C. Ainsworth, F. K. Sparrow & A. S. Sussman.<br />

New York: Academic Press, pp. 451 478.<br />

Drinkard, L. C., Nelson, G. E. & Sutter, R. P. (1982).<br />

Growth arrest: a prerequisite for sexual development<br />

in Phycomyces blakesleeanus. Experimental<br />

Mycology, 6, 52 59.<br />

Dubourdieu, D., Pucheu-Planté, B., Mercier, M. &<br />

Ribéreau-Gayon, P. (1978a). Structure, rôle et localisation<br />

du glucane exo-cellulaire sécrété par Botrytis<br />

cinerea dans la baie de Raisin. Comptes Rendus<br />

Hebdomadaires des Séances de l’Académie des Sciences, D<br />

287, 571 573.<br />

Dubourdieu, D., Fournet, B., Bertrand, A. & Ribéreau-<br />

Gayon, P. (1978b). Identification du glucane sécrété<br />

dans la baie de Raisin par Botrytis cinerea. Comptes<br />

Rendus Hebdomadaires des Séances de l’Académie des<br />

Sciences, D 286, 229 231.<br />

Dudding<strong>to</strong>n, C. L. (1955). <strong>Fungi</strong> that attack microscopic<br />

animals. Biological Reviews, 21, 377 439.<br />

Dudding<strong>to</strong>n, C. L. (1957). The Friendly <strong>Fungi</strong>. A New<br />

Approach <strong>to</strong> the Eelworm Problem. London: Faber and<br />

Faber.<br />

Dufour, E., Boulay, J., Rincheval, V. & Sainsard-Chanet,<br />

A. (2000). A causal link between respiration and<br />

senescence in Podospora anserina. Proceedings of the<br />

National Academy of Sciences of the USA, 97, 4138 4143.<br />

Dughi, R. (1956). La signification des appareils apicaux<br />

des asques de Gymnophysma et de Chlamydophysma.<br />

Comptes Rendus Hebdomadaires des Séances de l’Académie<br />

des Sciences de Paris, 243, 750 752.<br />

Duncan, E. G. & Galbraith, M. H. (1972). Post-meiotic<br />

events in Homobasidiomycetidae. Transactions of the<br />

British Mycological Society, 58, 387 392.<br />

Duncan, K. E. & Howard, R. J. (2000). Cy<strong>to</strong>logical<br />

analysis of wheat infection by the leaf blotch<br />

pathogen Mycosphaerella graminicola. Mycological<br />

Research, 104, 1074 1082.<br />

Duncan, R. A., Sullivan, R., Alderman, S. C., Spatafora,<br />

J. W. & White, J. F. (2002). An ergot adapted <strong>to</strong> the<br />

aquatic environment. Mycotaxon, 81, 11 25.<br />

Dunlap, J. C. (1999). Molecular bases for circadian<br />

clocks. Cell, 96, 271 290.<br />

Dunlap, J. C. & Loros, J. J. (2004). The Neurospora<br />

circadian system. Journal of Biological Rhythms, 19,<br />

414 424.<br />

Dunphy, G. B. & Nolan, R. A. (1982). Cellular immune<br />

response of spruce budworm larvae <strong>to</strong> En<strong>to</strong>mophthora<br />

aulicae and other test particles. Invertebrate Pathology,<br />

39, 81 92.<br />

Dupont, B., Crewe Brown, H. H., Westermann, K., et al.<br />

(2000). Mycoses in AIDS. Medical Mycology, 38,<br />

Supplement 1, 259 267.<br />

Durand, H., Clanet, M. & Tiraby, G. (1988). Genetic<br />

improvement of Trichoderma reesei for large scale<br />

cellulase production. Enzyme and Microbial Technology,<br />

10, 341 346.<br />

Dyal, S. D. & Narine, S. S. (2005). Implications for the<br />

use of Mortierella fungi in the industrial production<br />

of essential fatty acids. Food Research International, 38,<br />

445 467.<br />

Dyck, P. L. & Kerber, E. R. (1985). Resistance of the racespecific<br />

type. In The Cereal Rusts 2: Diseases,<br />

Distribution, Epidemiology, and Control, ed. A. P. Roelfs &<br />

W. R. Bushnell. Orlando: Academic Press,<br />

pp. 469 500.<br />

Dykstra, M. J. (1994). Ballis<strong>to</strong>sporic conidia in<br />

Basidiobolus ranarum: the influence of light and<br />

nutrition on the production of conidia and<br />

endospores (sporangiospores). Mycologia, 86,<br />

494 501.<br />

Dykstra, M. J. & Bradley-Kerr, B. (1994). The adhesive<br />

droplet of capilliconidia of Basidiobolus ranarum<br />

exhibits unique ultrastructural features. Mycologia,<br />

86, 336 342.<br />

Dylewski, D. P. (1990). Phylum Plasmodiophoromycota.<br />

In Handbook of Pro<strong>to</strong>ctista, ed. L. Margulis, J. O. Corliss,<br />

M. Melkonian & D. J. Chapman. Bos<strong>to</strong>n: Jones &<br />

Bartlett, pp. 399 416.<br />

Eadie, M. J. (2004). Ergot of rye the first specific for<br />

migraine. Journal of Clinical Neuroscience, 11, 4 7.<br />

Ea<strong>to</strong>n, R. A. & Hale, M. D. C. (1993). Wood, Decay, Pests and<br />

Prevention. London: Chapman & Hall.<br />

Eckstein, P. E., Krasichynska, N., Voth, D., et al.<br />

(2002). Development of PCR-based markers for a<br />

gene (Un8) conferring true loose smut resistance<br />

in barley. Canadian Journal of Plant Pathology, 24,<br />

46 53.<br />

Edelmann, R. E. & Klomparens, K. L. (1994). The ultrastructural<br />

development of sporangiospores in<br />

multispored sporangia of Zygorhynchus heterogamus<br />

with a hypothesis for sporangial wall dissolution.<br />

Mycologia, 86, 57 71.


REFERENCES<br />

731<br />

Edgar, J. A., Frahn, J. L., Cockrum, P. A. & Culvenor,<br />

C. C. J. (1986). Lupinosis. The chemistry and<br />

biochemistry of the phomopsins. In Myco<strong>to</strong>xins and<br />

Phy<strong>to</strong><strong>to</strong>xins, ed. P. S. Stern & R. Vleggaar. Amsterdam:<br />

Elsevier, pp. 169 86.<br />

Edlind, T. D. & Katiyar, S. K. (2004). The echinocandin<br />

‘target’ identified by cross-linking is a homolog of<br />

Pil1 and Lsp1, sphingolipid-dependent regula<strong>to</strong>rs of<br />

cell wall integrity signaling. Antimicrobial Agents and<br />

Chemotherapy, 48, 4491.<br />

Edwards, H. H. & Allen, P. J. (1970). A fine-structure<br />

study of the primary infection process of barley<br />

infected with Erysiphe graminis f. sp. hordei.<br />

Phy<strong>to</strong>pathology, 60, 1504 1509.<br />

Eggert, C., Temp, U. & Eriksson, K.-E. L. (1997). Laccase is<br />

essential for lignin degradation by the white-rot<br />

fungus Pycnoporus cinnabarinus. FEBS Letters, 407,<br />

89 92.<br />

Eisfelder, I. & Herschel, K. (1966). Agathomyia wankowiczi<br />

Schnabl, die ‘Zitzengallenfliege’ aus Ganoderma<br />

applanatum. Westfälische Pilzbriefe, 6, 5 10.<br />

El-Abyad, M. S. H. & Webster, J. (1968a). Studies on<br />

pyrophilous Discomycetes. I. Comparative physiological<br />

studies. Transactions of the British Mycological<br />

Society, 51, 353 367.<br />

El-Abyad, M. S. H. & Webster, J. (1968b). Studies on<br />

pyrophilous Discomycetes. II. Competition.<br />

Transactions of the British Mycological Society, 51,<br />

369 375.<br />

Elad, Y., Lifshitz, R. & Baker, R. (1985). Enzymatic<br />

activity of the mycoparasite Pythium nunn during<br />

interaction with host and non-host fungi.<br />

Physiological Plant Pathology, 27, 131 148.<br />

Elad, Y., Williamson, B., Tudzynski, P. & Delen, N., eds.<br />

(2004). Botrytis: Biology, Pathology and Control.<br />

Dordrecht: Kluwer.<br />

El-Ani, A. S. (1971). Chromosome numbers in the<br />

Hypocreales. II. Ascus development in Nectria cinnabarina.<br />

American Journal of Botany, 58, 56 60.<br />

Elix, J. A. (1996). Biochemistry and secondary metabolites.<br />

In Lichen Biology, ed. T. H. Nash. Cambridge:<br />

Cambridge University Press, pp. 154 180.<br />

Ellingboe, A. H. (1961). Somatic recombination in<br />

Puccinia graminis var. tritici. Phy<strong>to</strong>pathology, 51, 13 15.<br />

Elliott, C. G. (1994). Reproduction in <strong>Fungi</strong>. Genetic and<br />

Physiological Aspects. London: Chapman & Hall.<br />

Elliott, E. W. (1949). The swarm cells of myxomycetes.<br />

Mycologia, 41, 141 170.<br />

Elliott, T. J. (1985). The general biology of the mushroom.<br />

In The Biology and Technology of the Cultivated<br />

Mushroom, ed. P. B. Flegg, D. M. Spencer & D. A. Wood.<br />

Chichester: John Wiley & Sons, pp. 9 22.<br />

Ellis, D. H. & Pfeiffer, T. J. (1990a). Natural habitat of<br />

Cryp<strong>to</strong>coccus neoformans var. gattii. Journal of Clinical<br />

Microbiology, 28, 1642 1644.<br />

Ellis, D. H. & Pfeiffer, T. J. (1990b). Ecology, life cycle,<br />

and infectious propagule of Cryp<strong>to</strong>coccus neoformans.<br />

The Lancet, 336, 923 925.<br />

Ellis, M. B. (1971a). Dematiaceous Hyphomycetes. Kew:<br />

Commonwealth Mycological Institute.<br />

Ellis, M. B. (1971b). Porospores. In Taxonomy of <strong>Fungi</strong><br />

Imperfecti, ed. W. B. Kendrick. Toron<strong>to</strong>: University of<br />

Toron<strong>to</strong> Press, pp. 71 4.<br />

Ellis, M. B. (1976). More Dematiaceous Hyphomycetes. Kew:<br />

Commonwealth Mycological Institute.<br />

Ellis, M. B. & Ellis, J. P. (1990). <strong>Fungi</strong> Without Gills<br />

(Hymenomycetes and Gasteromycetes). London: Chapman<br />

& Hall.<br />

Ellis, M. B. & Ellis, J. P. (1998). Microfungi on Miscellaneous<br />

Substrates. Slough: Richmond Publishing.<br />

El-Shafie, A. K. & Webster, J. (1980). Ascospore liberation<br />

in Cochliobolus cymbopogonis. Transactions of the British<br />

Mycological Society, 75, 141 146.<br />

Embley, T. M., van der Giezen, M., Horner, D. S., Dyal,<br />

P. L. & Foster, P. (2002). Mi<strong>to</strong>chondria and hydrogenosomes<br />

are two forms of the same fundamental<br />

organelle. Philosophical Transactions of the Royal Society,<br />

B 358, 191 203.<br />

Emerson, R. (1941). An experimental study of the life<br />

cycles and taxonomy of Allomyces. Lloydia, 4, 77 144.<br />

Emerson, R. (1958). Mycological organization.<br />

Mycologia, 50, 589 621.<br />

Emerson, R. & Natvig, D. O. (1981). Adaptation of fungi<br />

<strong>to</strong> stagnant waters. In The Fungal Community. Its<br />

Organization and Role in the Ecosystem, ed. D. T.<br />

Wicklow & G. C. Carroll. New York: Marcel Dekker,<br />

pp. 109 128.<br />

Emerson, R. & Robertson, J. A. (1974). Two new<br />

members of the Blas<strong>to</strong>cladiaceae. I. Taxonomy,<br />

with evaluation of genera and inter-relationships<br />

in the family. American Journal of Botany, 61,<br />

303 317.<br />

Emerson, R. & Wes<strong>to</strong>n, W. H. (1967). Aqualinderella<br />

fermentans gen. et sp. nov., a phycomycete adapted<br />

<strong>to</strong> stagnant waters. I. Morphology and<br />

occurrence in nature. American Journal of Botany,<br />

54, 702 719.<br />

Emerson, R. & Wilson, C. M. (1954). Interspecific<br />

hybrids and the cy<strong>to</strong>genetics and cy<strong>to</strong>taxonomy of<br />

Eu-Allomyces. Mycologia, 46, 393 434.<br />

Emmons, C. W. (1955). Saprophytic sources of<br />

Cryp<strong>to</strong>coccus neoformans associated with the pigeon<br />

(Columba livia). American Journal of Hygiene, 62,<br />

227 232.


732 REFERENCES<br />

Engel, H. & Schneider, J. C. (1963). Die Umwandlung<br />

von Glykogen in Zucker in den Fruchtkörpern von<br />

Sphaerobolus stellatus (Tode) Pers., vor ihrem Abschuss.<br />

Berichte der Deutschen Botanischen Gesellschaft, 75,<br />

397 400.<br />

Englander, L. & Hull, R. J. (1980). Reciprocal transfer of<br />

nutrients between ericaceous plants and a Clavaria<br />

sp. New Phy<strong>to</strong>logist, 84, 661 667.<br />

Engler, M., Anke, T. & Sterner, O. (1998). Production of<br />

antibiotics by Collybia nivalis, Omphalotus olearius, a<br />

Favolaschia and a Pterula species on natural substrates.<br />

Zeitschrift für Naturforschung, 53c, 318 324.<br />

Epstein, L., Laccetti, L., Staples, R. C., Hoch, H. C. &<br />

Hoose, W. A. (1985). Extracellular proteins associated<br />

with induction of differentiation in bean rust<br />

uredospore germlings. Phy<strong>to</strong>pathology, 75,<br />

1073 1076.<br />

Eriksson, O. E. (1981). The families of bitunicate<br />

ascomycetes. Opera Botanica, 60, 1 220.<br />

Eriksson, O. E., Baral, H.-O., Currah, R. S., et al. (2003).<br />

Outline of Ascomycota. Myconet, 9, 1 89.<br />

Erwin, D. C. & Ribeiro, O. K. (1996). Phy<strong>to</strong>phthora Diseases<br />

Worldwide. St. Paul: APS Press.<br />

Erwin, D. C., Bartnicki-Garcia, S. & Tsao, P. H., eds.<br />

(1983). Phy<strong>to</strong>phthora. Its Biology, Taxonomy, Ecology, and<br />

Pathology. St. Paul: APS Press.<br />

Escobar, G., McCabe, D. E. & Harpel, C. W. (1976).<br />

Limnoperdon, a floating Gasteromycete isolated from<br />

marshes. Mycologia, 68, 874 880.<br />

Eslava, A. P. & Alvarez, M. I. (1987). Crosses. In<br />

Phycomyces, ed. E. Cerdá-Olmedo & E. D. Lipson. Cold<br />

Spring Harbor, New York: Cold Spring Harbor<br />

Labora<strong>to</strong>ry, pp. 361 365.<br />

Eslava, A. P., Alvarez, M. I. & Delbrück, M. (1975a).<br />

Meiosis in Phycomyces. Proceedings of the National<br />

Academy of Sciences of the USA, 72, 4076 4080.<br />

Eslava, A. P., Alvarez, M. I., Burke, P. V. & Delbrück, M.<br />

(1975b). Genetic recombination in sexual crosses in<br />

Phycomyces. Genetics, 80, 445 462.<br />

Esser, K. (1974). Podospora anserina. In Handbook of<br />

Genetics, ed. R. C. King. New York: Plenum Press,<br />

pp. 531 551.<br />

Esser, K. & Blaich, R. (1994). Heterogenic incompatibility<br />

in fungi. In The Mycota I: Growth, Differentiation<br />

and Sexuality, ed. J. G. H. Wessels & F. Meinhardt.<br />

Berlin: Springer-Verlag, pp. 211 232.<br />

Estey, R. H. & Tzean, S. S. (1976). Scanning electron<br />

microscopy of fungal nema<strong>to</strong>de-trapping devices.<br />

Transactions of the British Mycological Society, 66,<br />

520 522.<br />

Eucker, J., Sezer, O., Graf, B. & Possinger, K. (2001).<br />

Mucormycoses. Mycoses, 44, 253 260.<br />

Evans, G. H. & Cooke, R. C. (1982). Studies on<br />

Mucoralean mycoparasites. 3. Diffusible fac<strong>to</strong>rs from<br />

Mortierella vinacea Dixon-Stewart that direct germ<br />

tube growth in Pip<strong>to</strong>cephalis fimbriata Richardson &<br />

Leadbeater. New Phy<strong>to</strong>logist, 91, 245 253.<br />

Evans, G. H., Lewis, D. H. & Cooke, R. C. (1978). Studies<br />

on Mucoralean mycoparasites. 2. Persistent yeastphase<br />

growth of Mycotypha microspora Fenner when<br />

infected by Pip<strong>to</strong>cephalis fimbriata Richardson &<br />

Leadbeater. New Phy<strong>to</strong>logist, 81, 629 635.<br />

Evans, H. C. (1989). Mycopathogens of insects of epigeal<br />

and aerial habitats. In Insect Fungus Interactions, ed.<br />

N. Wilding, N. M. Collins, P. M. Hammond & J. F.<br />

Webber. London: Academic Press, pp. 205 238.<br />

Evans, H. C. (2003). Use of Clavicipitalean fungi for the<br />

biological control of arthropod pests. In<br />

Clavicipitalean <strong>Fungi</strong>, ed. F. White, C. W. Bacon, N. L.<br />

Hywel-Jones & J. W. Spatafora. New York: Marcel<br />

Dekker Inc., pp. 517 548.<br />

Evans, H. C. & Samson, R. A. (1982). Cordyceps species<br />

and their anamorphs pathogenic on ants<br />

(Formicidae) in tropical forest ecosystems. I. The<br />

Cephalotes (Myrmicinae) complex. Transactions of the<br />

British Mycological Society, 79, 431 453.<br />

Evans, H. C. & Samson, R. A. (1984). Cordyceps species<br />

and their anamorphs pathogenic on ants<br />

(Formicidae) in tropical forest ecosystems. II. The<br />

Campanotus (Formicinae) complex. Transactions of the<br />

British Mycological Society, 82, 127 150.<br />

Evans, H. C., Holmes, K. A., Phillips, W. & Wilkinson,<br />

M. J. (2002). What’s in a name: Crinipellis, the final<br />

resting place for the frosty pod rot pathogen of<br />

cocoa. Mycologist, 16, 148 152.<br />

Evans, H. C., Holmes, K. A. & Reid, A. P. (2003).<br />

Phylogeny of the frosty pod rot pathogen of cocoa.<br />

Plant Pathology, 52, 476 485.<br />

Evans, H. J. (1959). Nuclear behaviour in the cultivated<br />

mushroom. Chromosoma, 10, 411 419.<br />

Evans, R. G. (1965). Sporobolomyces as a cause of<br />

respira<strong>to</strong>ry allergy. Acta Allergologica, 20, 197 205.<br />

Eversmeyer, M. G. & Kramer, C. L. (2000). Epidemiology<br />

of wheat leaf and stem rust in the Central Great<br />

Plains of the USA. Annual Review of Phy<strong>to</strong>pathology, 38,<br />

491 513.<br />

Eyal, Z. (1999). The Sep<strong>to</strong>ria tritici and Stagonospora<br />

nodorum blotch diseases of wheat. European Journal of<br />

Plant Pathology, 105, 629 641.<br />

Fahselt, D. (1994). Secondary biochemistry of lichens.<br />

Symbiosis, 16, 117 165.<br />

Fahselt, D. (1996). Individuals, populations and population<br />

ecology. In Lichen Biology, ed. T. H. Nash.<br />

Cambridge: Cambridge University Press, pp. 181 98.


REFERENCES<br />

733<br />

Falk, S. P., Gadoury, D. M., Cortesi, P., Pearson, R. C. &<br />

Seem, R. C. (1995). Parasitism of Uncinula neca<strong>to</strong>r<br />

cleis<strong>to</strong>thecia by the mycoparasite Ampelomyces quisqualis.<br />

Phy<strong>to</strong>pathology, 85, 794 800.<br />

Fan, K. W., Vrijmoed, L. L. P. & Jones, E. B. G. (2002).<br />

Zoospore chemotaxis of mangrove thraus<strong>to</strong>chytrids<br />

from Hong Kong. Mycologia, 94, 569 578.<br />

Fang, J. G. & Tsao, P. H. (1995). Evaluation of Pythium<br />

nunn as a potential biocontrol agent against<br />

Phy<strong>to</strong>phthora root rot of azalea and sweet orange.<br />

Phy<strong>to</strong>pathology, 85, 29 36.<br />

Faretra, F., An<strong>to</strong>nacci, E. & Pollastro, S. (1988). Sexual<br />

behaviour and mating system of Botryotinia fuckeliana,<br />

teleomorph of Botrytis cinerea. Journal of General<br />

Microbiology, 134, 2543 2550.<br />

Farr, D. F., Bills, G. F., Chamuris, G. P. & Rossman, A. Y.<br />

(1989). <strong>Fungi</strong> on Plants and Plant Products in the United<br />

States. St. Paul: APS Press.<br />

Fasolo-Bonfante, P. & Brunel, A. (1972). Caryological<br />

features in a mycorrhizal fungus: Tuber. Allionia, 18,<br />

5 11.<br />

Fassi, B., Fontana, A. & Trappe, J. M. (1969).<br />

Ec<strong>to</strong>mycorrhizae formed by Endogone lactiflua with<br />

species of Pinus and Pseudotsuga. Mycologia, 61,<br />

412 414.<br />

Fatehi, J., Bridge, P. D. & Punithalingam, E. (2003).<br />

Molecular relatedness within the ‘Ascochyta pinodescomplex’.<br />

Mycopathologia, 156, 317 327.<br />

Faulstich, H. & Zilker, T. R. (1994). Ama<strong>to</strong>xins. In<br />

Handbook of Mushroom Poisoning. Diagnosis and<br />

Treatment, 2nd edn, ed. D. G. Spoerke & B. H. Rumach.<br />

Boca Ra<strong>to</strong>n: CRC Press, pp. 233 264.<br />

Federici, B. A. (1977). Differential pigmentation in the<br />

sexual phase of Coelomomyces. Nature, 267, 514 515.<br />

Feldbrügge, M., Kämper, J., Steinberg, G. & Kahmann,<br />

R. (2004). Regulation of mating and pathogenic<br />

development in Ustilago maydis. Current Opinion in<br />

Microbiology, 7, 666 672.<br />

Fell, J. W. & Blatt, G. M. (1999). Separation of strains of<br />

the yeasts Xanthophyllomyces dendrorhous and Phaffia<br />

rhodozyma based on rDNA IGS and ITS sequence<br />

analysis. Journal of Industrial Microbiology and<br />

Biotechnology, 23, 677 681.<br />

Fell, J. W. & Statzell-Tallman, A. (1981). Heterothallism<br />

in the basidiomyce<strong>to</strong>us genus Sporidiobolus. Current<br />

Microbiology, 5, 77 82.<br />

Fell, J. W. & Statzell-Tallman, A. (1998). Rhodosporidium<br />

Banno. In The Yeasts. A Taxonomic Study, 4th edn, ed.<br />

C. P. Kurtzman & J. W. Fell. Amsterdam: Elsevier,<br />

pp. 678 692.<br />

Fell, J. W., Boekhout, T., Fonseca, A., Scorzetti, G. &<br />

Statzell-Tallman, A. (2000). Biodiversity and<br />

systematics of basidiomyce<strong>to</strong>us yeasts as determined<br />

by large-subunit rDNA D1/D2 domain sequence<br />

analysis. International Journal of Systematic and<br />

Evolutionary Microbiology, 50, 1351 1371.<br />

Fell, J. W., Boekhout, T., Fonseca, A. & Sampaio, J. P.<br />

(2001). Basidiomyce<strong>to</strong>us yeasts. In The Mycota VIIB:<br />

Systematics and Evolution, ed. D. J. McLaughlin, E. G.<br />

McLaughlin & P. A. Lemke. Berlin: Springer-Verlag,<br />

pp. 3 35.<br />

Fiasson, J. L. & Petersen, R. H. (1973). Carotenes in the<br />

fungus Clathrus ruber (Gasteromycetes). Mycologia, 65,<br />

201 203.<br />

Fiddy, C. & Trinci, A. P. J. (1976). Mi<strong>to</strong>sis, septation,<br />

branching and the duplication cycle in<br />

Aspergillus nidulans. Journal of General Microbiology,<br />

97, 169 184.<br />

Field, J. I. & Webster, J. (1977). Traps of predacious fungi<br />

attract nema<strong>to</strong>des. Transactions of the British<br />

Mycological Society, 68, 467 470.<br />

Field, J. I. & Webster, J. (1983). Anaerobic survival of<br />

aquatic fungi. Transactions of the British Mycological<br />

Society, 81, 365 369.<br />

Field, J. I. & Webster, J. (1985). Effects of sulphide on<br />

survival of aero-aquatic and aquatic hyphomycetes.<br />

Transactions of the British Mycological Society, 85,<br />

193 199.<br />

Filtenborg, O., Frisvad, J. C. & Samson, R. A. (2002).<br />

Specific association of fungi <strong>to</strong> foods and influence<br />

of physical environmental fac<strong>to</strong>rs. In <strong>Introduction</strong> <strong>to</strong><br />

Food- and Airborne <strong>Fungi</strong>, 6th edn, ed. R. A. Samson,<br />

E. S. Hoekstra, J. C. Frisvad & O. Filtenborg.<br />

Utrecht: Centraalbureau voor Schimmelcultures,<br />

pp. 306 20.<br />

Fincham, J. R. S. & Day, P. R. (1971). Fungal Genetics, 3rd<br />

edn. Oxford: Blackwell Scientific.<br />

Finger, F. P. & Novick, P. (1998). Spatial regulation of<br />

exocy<strong>to</strong>sis: lessons from yeast. Journal of Cell Biology,<br />

142, 609 612.<br />

Fischer, A. (1892). Die Pilze. IV Abt. Phycomycetes. In<br />

Kryp<strong>to</strong>gamenflora von Deutschland, Oesterreich und<br />

der Schweiz, ed. L. Rabenhorst. Leipzig: Eduard<br />

Kummer.<br />

Fischer, F. G. & Werner, G. (1958). Die Chemotaxis der<br />

Schwärmsporen von Wasserpilzen<br />

(Saprolegniaceen). Hoppe-Seyler’s Zeitschrift für<br />

Physiologische Chemie, 310, 92 96.<br />

Fischer, G. W. & Hol<strong>to</strong>n, C. S. (1957). Biology and Control<br />

of the Smut <strong>Fungi</strong>. New York: Ronald Press Company.<br />

Fischer, M., Cox, J., Davis, D. J., et al. (2004). New<br />

information on the mechanism of forcible ascospore<br />

discharge from Ascobolus immersus. Fungal Genetics and<br />

Biology, 41, 698 707.


734 REFERENCES<br />

Fischer-Par<strong>to</strong>n, S., Par<strong>to</strong>n, R. M., Hickey, P. C., et al. (2000).<br />

Confocal microscopy of FM4 64 as a <strong>to</strong>ol for<br />

analysing endocy<strong>to</strong>sis and vesicle trafficking in living<br />

fungal hyphae. Journal of Microscopy, 198, 246 259.<br />

Fisher, K. E., Lowry, D. S. & Roberson, R. W. (2000).<br />

Cy<strong>to</strong>plasmic cleavage in living sporangia of Allomyces<br />

macrogynus. Journal of Microscopy, 198, 260 269.<br />

Fisher, M. C., Koenig, G. L., White, T. J. & Taylor, J. T.<br />

(2002). Molecular and phenotypic description of<br />

Coccidioides posadasii sp. nov., previously recognized<br />

as the non-California population of Coccidioides<br />

immitis. Mycologia, 94, 73 84.<br />

Fisher, P. J. (1978). Survival of aero-aquatic fungi on<br />

land. Transactions of the British Mycological Society, 71,<br />

419 423.<br />

Fisher, P. J. & Webster, J. (1978). Sporulation of aeroaquatic<br />

fungi under different gas regimes in light<br />

and darkness. Transactions of the British Mycological<br />

Society, 71, 465 468.<br />

Fisher, P. J. & Webster, J. (1979). Effect of oxygen and<br />

carbon dioxide on growth of four aero-aquatic<br />

hyphomycetes. Transactions of the British Mycological<br />

Society, 72, 57 61.<br />

Fisher, P. J., Webster, J. & Petrini, O. (1991). Aquatic<br />

hyphomycetes and other fungi in living aquatic and<br />

terrestrial roots of Alnus glutinosa. Mycological<br />

Research, 95, 543 547.<br />

Fitt, B. D. L., Goulds, A. & Polley, R. W. (1988). Eyespot<br />

(Pseudocercosporella herpotrichoides) epidemiology in<br />

relation <strong>to</strong> prediction of disease severity and yield<br />

loss in winter wheat: a review. Plant Pathology, 37,<br />

311 328.<br />

Flaishman, M. A. & Kollatukudy, P. E. (1994). Timing of<br />

fungal invasion using host’s ripening hormone as a<br />

signal. Proceedings of the National Academy of Sciences of<br />

the USA, 91, 6579 6583.<br />

Fleet, G. H. (1990). Food spoilage yeasts. In Yeast<br />

Technology, ed. J. F. T. Spencer & D. M. Spencer. Berlin:<br />

Springer-Verlag, pp. 124 166.<br />

Flegg, P. B. & Wood, D. A. (1985). Growth and fruiting.<br />

In The Biology and Technology of the Cultivated Mushroom,<br />

ed. P. B. Flegg, D. M. Spencer & D. A. Wood.<br />

Chichester: John Wiley & Sons, pp. 141 177.<br />

Flegler, S. L. (1984). An improved method for production<br />

of Sphaerobolus fruit bodies in culture. Mycologia,<br />

76, 944 946.<br />

Fleming, A. (1929). On the antibacterial action of<br />

cultures of Penicillium, with a special reference <strong>to</strong><br />

their use in the isolation of B. influenzae. British<br />

Journal of Experimental Pathology, 10, 226 236.<br />

Fleming, A. (1944). The discovery of penicillin. British<br />

Medical Bulletin, 2, 4 5.<br />

Fletcher, H. J. (1969). The development and tropisms of<br />

Pilaira anomala. Transactions of the British Mycological<br />

Society, 53, 130 132.<br />

Fletcher, H. J. (1973). The sporangiophore of Pilaira<br />

species. Transactions of the British Mycological Society, 61,<br />

553 568.<br />

Fletcher, J. (1972). Fine structure of developing merosporangia<br />

and sporangiospores of Syncephalastrum<br />

racemosum. Archiv für Mikrobiologie, 87, 269 284.<br />

Fletcher, J. (1973a). The distribution of cy<strong>to</strong>plasmic<br />

vesicles, multivesicular bodies and paramural bodies<br />

in elongating sporangiophores and swelling sporangia<br />

of Thamnidium elegans Link. Annals of Botany, 37,<br />

955 961.<br />

Fletcher, J. (1973b). Ultrastructural changes associated<br />

with spore formation in sporangia and sporangiola<br />

of Thamnidium elegans Link. Annals of Botany, 37,<br />

963 972.<br />

Fletcher, L. R. & Harvey, I. C. (1981). An association of a<br />

Lolium endophyte with ryegrass staggers. New Zealand<br />

Veterinary Journal, 29, 185 186.<br />

Flieger, M., Wurst, M. & Shelby, R. (1997). Ergot<br />

alkaloids sources, structures and analytical<br />

methods. Folia Microbiologica, 42, 3 30.<br />

Flor, H. H. (1955). Host parasite interaction in flax rust<br />

its genetics and other implications. Phy<strong>to</strong>pathology,<br />

45, 680 685.<br />

Flor, H. H. (1971). Current status of the gene-for-gene<br />

concept. Annual Review of Phy<strong>to</strong>pathology, 9, 275 296.<br />

Flores, R., Dederichs, A., Cerdá-Olmedo, E. & Hertel, R.<br />

(1999). Flavin-binding sites in Phycomyces. Plant<br />

Biology, 1, 645 655.<br />

Fokkema, N. J. & van der Meulen, F. (1976). Antagonism<br />

of yeastlike phyllosphere fungi against Sep<strong>to</strong>ria<br />

nodorum on wheat leaves. Netherlands Journal of Plant<br />

Pathology, 82, 13 16.<br />

Foley, M. F. & Deacon, J. W. (1985). Isolation of Pythium<br />

oligandrum and other necrotrophic mycoparasites<br />

from soil. Transactions of the British Mycological Society,<br />

85, 631 639.<br />

Foley, M. F. & Deacon, J. W. (1986a). Physiological<br />

differences between mycoparasitic and plant-pathogenic<br />

Pythium spp. Transactions of the British Mycological<br />

Society, 86, 225 231.<br />

Foley, M. F. & Deacon, J. W. (1986b). Susceptibility of<br />

Pythium spp. and other soil fungi <strong>to</strong> antagonism by<br />

the mycoparasite Pythium oligandrum. Soil Biology and<br />

Biochemistry, 18, 91 95.<br />

Fontaine, T., Hartland, R. P., Beauvais, A., Diaquin, M. &<br />

Latgé, J.-P. (1997). Purification and characterization<br />

of an endo-1,3-b-glucanase from Aspergillus fumigatus.<br />

European Journal of Biochemistry, 243, 315 321.


REFERENCES<br />

735<br />

Foster, S. J. & Fitt, B. D. L. (2004). Isolation and characterisation<br />

of the mating-type (MAT) locus from<br />

Rhynchosporium secalis. Current Genetics, 44, 277 286.<br />

Fox, R. D. & Wong, G. J. (1990). Homothallism and<br />

heterothallism in Tremella fuciformis. Canadian Journal<br />

of Botany, 68, 107 111.<br />

Fox, R. T. V., ed. (2000). Armillaria Root Rot: Biology and<br />

Control of Honey Fungus. Andover, UK: Intercept Ltd.<br />

Franke, J. & Kessin, R. (1977). A defined minimal<br />

medium for axenic strains of Dictyostelium discoideum.<br />

Proceedings of the National Academy of Sciences of the USA,<br />

74, 2157 2161.<br />

Frankland, J. C. (1984). Autecology and the mycelium of<br />

a woodland litter decomposer. In The Ecology and<br />

Physiology of the Fungal Mycelium, ed. D. H. Jennings &<br />

A. D. M. Rayner. Cambridge: Cambridge University<br />

Press, pp. 241 260.<br />

Franzot, S. P., Salkin, I. F. & Casadevall, A. (1999).<br />

Cryp<strong>to</strong>coccus neoformans var. grubii: separate varietal<br />

status for Cryp<strong>to</strong>coccus neoformans serotype A isolates.<br />

Journal of Clinical Microbiology, 37, 838 840.<br />

Fraser, J. A. & Heitman, J. (2004). Evolution of fungal sex<br />

chromosomes. Molecular Microbiology, 51, 299 306.<br />

Fraymouth, J. (1956). Haus<strong>to</strong>ria of the Peronosporales.<br />

Transactions of the British Mycological Society, 39,79 107.<br />

Frazer, L. N. (1996). Control of growth and patterning in<br />

the fungal fruiting structure. A case for the<br />

involvement of hormones. In Patterns in Fungal<br />

Development, ed. S.-W. Chiu & D. Moore. Cambridge:<br />

Cambridge University Press, pp. 156 181.<br />

Frederick, L. (1990). Phylum plasmodial slime molds,<br />

class Myxomycota. In Handbook of Pro<strong>to</strong>ctista, ed. L.<br />

Margulis, J. O. Corliss, M. Melkonian & D. J. Chapman.<br />

Bos<strong>to</strong>n: Jones & Bartlett, pp. 467 483.<br />

Frederick, L., Uecker, F. A. & Benjamin, C. R. (1969). A<br />

new species of Neurospora from the soil of West<br />

Pakistan. Mycologia, 61, 1077 1084.<br />

Frederickson, D. E. & Mantle, P. G. (1989). Secondary<br />

conidition of Sphacelia sorghi on sorghum, a novel<br />

fac<strong>to</strong>r in the epidemiology of the ergot disease.<br />

Mycological Research, 93, 497 502.<br />

Frederickson, D. E., Mantle, P. G. & De Milliano, W. A. J.<br />

(1991). Claviceps africana sp. nov.: the distinctive ergot<br />

pathogen of sorghum in Africa. Mycological Research,<br />

95, 1101 1107.<br />

Freeman, S. & Rodriguez, R. J. (1993). Genetic conversion<br />

of a fungal pathogen <strong>to</strong> a non-pathogenic,<br />

endophytic mutualist. Science, 260, 75 78.<br />

Frey, K. J., Browning, J. A. & Simons, M. D. (1973).<br />

Management of host resistance genes <strong>to</strong> control<br />

diseases. Zeitschrift für Pflanzenkrankheiten und<br />

Pflanzenschutz, 80, 160 180.<br />

Freytag, S. & Mendgen, K. (1991). Carbohydrates on the<br />

surface of urediniospore- and basidiospore-derived<br />

infection structures of heteroecious and au<strong>to</strong>ecious<br />

rust fungi. New Phy<strong>to</strong>logist, 119, 527 534.<br />

Friend, J. (1991). The biochemistry and cell biology of<br />

interaction. In Phy<strong>to</strong>phthora infestans, the Cause of Late<br />

Blight of Pota<strong>to</strong>, ed. D. S. Ingram & P. H. Williams.<br />

London: Academic Press, pp. 85 129.<br />

Friman, E., Olsson, S. & Nordbring-Hertz, B. (1985).<br />

Heavy trap formation by Arthrobotrys oligospora in<br />

liquid culture. FEMS Microbiology Ecology, 31, 17 21.<br />

Frisvad, J. C. & Samson, R. A. (2004). Polyphasic taxonomy<br />

of Penicillium subgenus Penicillium. A guide <strong>to</strong><br />

identification of food and air-borne terverticillate<br />

Penicillia and their myco<strong>to</strong>xins. Studies in Mycology,<br />

49, 1 173.<br />

Frisvad, J. C., Smedsgaard, J., Larsen, T. O. & Samson,<br />

R. A. (2004). Myco<strong>to</strong>xins, drugs and other extrolites<br />

produced by species in Penicillium subgenus<br />

Penicillium. Studies in Mycology, 49, 201 242.<br />

Fry, W. E. & Goodwin, S. B. (1997). Resurgence of the<br />

Irish pota<strong>to</strong> famine fungus. BioScience, 47, 363 371.<br />

Fry, W. E., Goodwin, S. B., Dyer, A. T., et al. (1993).<br />

His<strong>to</strong>rical and recent migrations of Phy<strong>to</strong>phthora<br />

infestans: chronology, pathways, and implications.<br />

Plant Disease, 77, 653 661.<br />

Fuentes-Dávila, G., Goates, B. J., Thomas, P., Nielsen, J. &<br />

Ballantyne, B. (2002). Smut diseases. In Bread Wheat.<br />

Improvement and Production, ed. B. C. Curtis, S. Rajaram<br />

&H.Gómez Macpherson. Rome: Food and Agriculture<br />

Organization of the United Nations, pp. 251 71.<br />

Fuhrer, B. (2005). A Field Guide <strong>to</strong> Australian <strong>Fungi</strong>.<br />

Melbourne: Bloomings Books Pty Ltd.<br />

Fukshansky, L. (1993). Intracellular processing of a<br />

spatially non-uniform stimulus: case study of<br />

pho<strong>to</strong>tropism in Phycomyces. Journal of Pho<strong>to</strong>chemistry<br />

and Pho<strong>to</strong>biology B: Biology, 19, 161 186.<br />

Fuller, J. C. (1969). The Day of St Anthony’s Fire. New York:<br />

Macmillan.<br />

Fuller, M. S. (1976). The zoospore, hallmark of aquatic<br />

fungi. Mycologia, 69, 1 20.<br />

Fuller, M. S. (1990). Phylum Hyphochytriomycota. In<br />

Handbook of Pro<strong>to</strong>ctista, ed. L. Margulis, J. O. Corliss, M.<br />

Melkonian & D. J. Chapman. Bos<strong>to</strong>n: Jones & Bartlett,<br />

pp. 380 7.<br />

Fuller, M. S. (2001). Hyphochytriomycota. In The Mycota<br />

VIIA: Systematics and Evolution, ed. D. J. McLaughlin,<br />

E. G. McLaughlin & P. A. Lemke. Berlin: Springer-<br />

Verlag, pp. 73 80.<br />

Fuller, M. S. & Clay, R. P. (1993). Observations on<br />

Gonapodya in pure culture: growth, development and<br />

cell wall characterization. Mycologia, 85, 38 45.


736 REFERENCES<br />

Fuller, M. S. & Olson, L. W. (1971). The zoospore of<br />

Allomyces. Journal of General Microbiology, 66, 171 183.<br />

Fuller, M. S. & Rakatansky, R. M. (1966). A preliminary<br />

study of the carotenoids in Acrasis rosea. Canadian<br />

Journal of Botany, 44, 269 274.<br />

Furch, B. & Gooday, G. W. (1978). Sporopollenin in<br />

Phycomyces blakesleeanus. Transactions of the British<br />

Mycological Society, 70, 307 309.<br />

Furtado, S. (1965). Relation of microstructures <strong>to</strong> the<br />

taxonomy of the Ganoderma<strong>to</strong>ideae (Polyporaceae).<br />

Mycologia, 57, 588 611.<br />

Gadd, G. M., ed. (2001). <strong>Fungi</strong> in Bioremediation.<br />

Cambridge: Cambridge University Press.<br />

Gadoury, D. M. & Pearson, R. C. (1988). Initiation,<br />

development, dispersal, and survival of cleis<strong>to</strong>thecia<br />

of Uncinula neca<strong>to</strong>r in New York vineyards.<br />

Phy<strong>to</strong>pathology, 78, 1413 1421.<br />

Gage, M. J., Bruenn, J., Fisher, M., Sanders, D. & Smith,<br />

T. J. (2001). KP4 fungal <strong>to</strong>xin inhibits growth in<br />

Ustilago maydis by blocking calcium uptake. Molecular<br />

Microbiology, 41, 775 785.<br />

Gage, M. J., Rane, S. G., Hockerman, G. H. & Smith, T. J.<br />

(2002). The virally encoded fungal <strong>to</strong>xin KP4 specifically<br />

blocks L-type voltage-gated calcium channels.<br />

Molecular Pharmacology, 61, 936 944.<br />

Gall, A. M. & Elliot, C. G. (1985). Control of sexual<br />

reproduction in Pythium sylvaticum. Transactions of the<br />

British Mycological Society, 84, 629 636.<br />

Galland, P. & Tölle, N. (2003). Light-induced fluorescence<br />

changes in Phycomyces: evidence for blue light-recep<strong>to</strong>r<br />

associated flavo-semiquinones. Planta, 217, 971 982.<br />

Gammie, A. E., Brizzio, V. & Rose, M. D. (1998). Distinct<br />

morphological phenotypes of cell fusion mutants.<br />

Molecular Biology of the Cell, 9, 1395 1410.<br />

Gamow, R. I., Ruiz-Herrera, J. & Fischer, E. P. (1987). The<br />

cell wall of Phycomyces. In Phycomyces, ed. E. Cerdá-<br />

Olmedo & E. D. Lipson. Cold Spring Harbor, New<br />

York: Cold Spring Harbor Labora<strong>to</strong>ry, pp. 223 246.<br />

Gams, W. (1977). A key <strong>to</strong> the species of Mortierella.<br />

Persoonia, 9, 381 391.<br />

Gams, W. & Bissett, J. (1998). Morphology and identification.<br />

In Trichoderma and Gliocladium, I: Basic<br />

Biology, Taxonomy and Genetics, ed. C. P. Kubicek & G. F.<br />

Harman. London: Taylor & Francis, pp. 3 34.<br />

Gams, W. & Zare, R. (2003). A taxonomic review of the<br />

Clavicipitaceous anamorphs parasitizing nema<strong>to</strong>des<br />

and other microinvertebrates. In Clavicipitalean <strong>Fungi</strong>,<br />

ed. J. F. White, C. W. Bacon, N. L. Hywel-Jones & J. W.<br />

Spatafora. New York: Marcel Dekker Inc., pp. 17 73.<br />

Gamundí, I. J. (1991). Review of recent advances in the<br />

knowledge of the Cyttariales. Systema Ascomycetum,<br />

10, 69 77.<br />

Garbaye, G. (1994). Helper bacteria: a new dimension <strong>to</strong><br />

the mycorrhizal symbiosis. New Phy<strong>to</strong>logist, 128,<br />

197 210.<br />

Garber, E. D. & Ruddat, M. (2002). Transmission<br />

genetics of Microbotryum violaceum (Ustilago violacea): a<br />

case his<strong>to</strong>ry. Advances in Applied Microbiology, 51,<br />

107 127.<br />

Garber, R. C. & Aist, J. R. (1979). The ultrastructure of<br />

meiosis in Plasmodiophora brassicae<br />

(Plasmodiophorales). Canadian Journal of Botany, 57,<br />

2509 2518.<br />

García-Pajón, C. M. & Collado, I. G. (2003). Secondary<br />

metabolites isolated from Colle<strong>to</strong>trichum species.<br />

Natural Product Reports, 20, 426 431.<br />

Gardiner, R. B. & Day, A. W. (1988). Surface proteinaceous<br />

fibrils (fimbriae) on filamen<strong>to</strong>us fungi.<br />

Canadian Journal of Botany, 66, 2474 2484.<br />

Gargas, A. & Taylor, J. W. (1995). Phylogeny of discomycetes<br />

and early radiations of the apothecial<br />

Ascomycotina inferred from SSU rDNA sequence<br />

data. Experimental Mycology, 19, 7 15.<br />

Gargas, A., DePriest, P. T., Grube, M. & Tehler, A. (1995).<br />

Multiple origins of lichen symbioses in fungi<br />

suggested by SSU rDNA phylogeny. Science, 268,<br />

1492 1495.<br />

Garrett, S. D. (1953). Rhizomorph behaviour in<br />

Armillaria mellea (Vahl) Quél. I. Fac<strong>to</strong>rs controlling<br />

rhizomorph initiation by Armillaria mellea in pure<br />

culture. Annals of Botany N.S., 17, 63 79.<br />

Garrett, S. D. (1970). Pathogenic Root-Infecting <strong>Fungi</strong>.<br />

Cambridge: Cambridge University Press.<br />

Garrett, R. G. & Tomlinson, J. A. (1967). Isolate differences<br />

in Olpidium brassicae. Transactions of the British<br />

Mycological Society, 50, 429 435.<br />

Garrill, A. (1995). Transport. In The Growing Fungus, ed.<br />

N. A. R. Gow & G. M. Gadd. London: Chapman & Hall,<br />

pp. 163 181.<br />

Garrill, A., Jackson, S. L., Lew, R. R. & Heath, I. B. (1993).<br />

Ion channel activity and tip growth: tip localised<br />

stretch-activated channels generate an essential<br />

Ca 2þ gradient in the oomycete Saprolegnia ferax.<br />

European Journal of Cell Biology, 60, 358 365.<br />

Garrison, R. G. & Boyd, K. S. (1973). Dimorphism of<br />

Penicillium marneffei as observed by electron microscopy.<br />

Canadian Journal of Microbiology, 19,<br />

1305 1309.<br />

Gauger, W. L. (1961). The germination of zygospores of<br />

Rhizopus s<strong>to</strong>lonifer. American Journal of Botany, 48,<br />

427 429.<br />

Gauger, W. L. (1966). Sexuality in an azygosporic strain<br />

of Mucor hiemalis. I. Breakdown of the azygosporic<br />

component. American Journal of Botany, 53, 751 755.


REFERENCES<br />

737<br />

Gauger, W. L. (1975). Further studies on sexuality in<br />

azygosporic strains of Mucor hiemalis. Transactions of<br />

the British Mycological Society, 64, 113 118.<br />

Gäumann, E. (1959). Die Rostpilze Mitteleuropas. Bern:<br />

Büchler & Co.<br />

Gauriloff, L. P. & Fuller, M. S. (1987). Rhizophydium<br />

sphaerocarpum. InZoosporic <strong>Fungi</strong> in Teaching and<br />

Research, ed. M. S. Fuller & A. Jaworski. Athens, USA:<br />

Southeastern Publishing Corporation, pp. 18 19.<br />

Gay, J. L., Salzberg, A. & Woods, A. M. (1987). Dynamic<br />

experimental evidence for the plasmamembrane<br />

ATPase domain hypothesis of haus<strong>to</strong>rial transport<br />

and for ionic coupling of the haus<strong>to</strong>rium of Erysiphe<br />

graminis <strong>to</strong> the host cell (Hordeum vulgare). New<br />

Phy<strong>to</strong>logist, 107, 541 548.<br />

Gay, L., Chanzy, H., Bulone, V., Girard, V. & Fèvre, M.<br />

(1993). Synthesis in vitro of crystalline chitin by a<br />

solubilized enzyme from the cellulosic fungus<br />

Saprolegnia monoica. Journal of General Microbiology,<br />

139, 2117 2122.<br />

Gehrig, H., Schüssler, A. & Kluge, M. (1996). Geosiphon<br />

pyriforme, a fungus forming endocy<strong>to</strong>biosis with<br />

Nos<strong>to</strong>c (Cyanobacteria) is an ancestral member of the<br />

Glomales: evidence by SSU rRNA analysis. Journal of<br />

Molecular Evolution, 43, 71 81.<br />

Geiser, D. M. & LoBuglio, K. F. (2001). The monophyletic<br />

Plec<strong>to</strong>mycetes: Ascosphaerales, Onygenales,<br />

Eurotiales. In The Mycota VIIA: Systematics and<br />

Evolution, ed. D. J. McLaughlin, E. G. McLaughlin &<br />

P. A. Lemke. Berlin: Springer-Verlag, pp. 201 19.<br />

Geiser, D. M., Timberlake, W. E. & Arnold, M. L. (1996).<br />

Loss of meiosis in Aspergillus. Molecular Biology and<br />

Evolution, 13, 809 817.<br />

Geiser, D. M., Dorner, J. W., Horn, B. W. & Taylor, J. W.<br />

(2000). The phylogenetics of myco<strong>to</strong>xin and sclerotium<br />

production in Aspergillus flavus and<br />

Aspergillus oryzae. Fungal Genetics and Biology, 31,<br />

169 179.<br />

Geiser, D. M., Jiménez-Gasco, M. M., Kang, S., et al.<br />

(2004). FUSARIUM-ID v. 1.0: a DNA sequence database<br />

for identifying Fusarium. European Journal of Plant<br />

Pathology, 110, 473 479.<br />

Gemmill, T. R. & Trimble, R. B. (1999). Overview of N-<br />

and O-linked oligosaccharide structures found in<br />

various yeast species. Biochimica et Biophysica Acta,<br />

1426, 227 237.<br />

Georgiou, C. D. & Petropoulou, K. P. (2001). Role of<br />

erythroascorbate and ascorbate in sclerotial differentiation<br />

in Sclerotinia sclerotiorum. Mycological<br />

Research, 105, 1364 1370.<br />

Georgiou, C. D., Tairis, N. & Polycratis, A. (2001).<br />

Production of b-carotene by Sclerotinia sclerotiorum<br />

and its role in sclerotium differentiation. Mycological<br />

Research, 105, 1100 1115.<br />

Georgopapadakou, N. H. (1998). Antifungals: mechanism<br />

of action and resistance, established and novel<br />

drugs. Current Opinion in Microbiology, 1, 547 557.<br />

Gerdemann, J. W. & Nicolson, T. H. (1963). Spores of<br />

mycorrhizal Endogone species extracted by wet<br />

sieving and decanting. Transactions of the British<br />

Mycological Society, 46, 235 244.<br />

Gerlach, W. & Nirenberg, H. I. (1982). The Genus Fusarium<br />

a Pic<strong>to</strong>rial Atlas. Berlin-Dahlem: Biologische<br />

Bundesanstalt für Land- und Forstwirtschaft.<br />

Gerlagh, M., Goossen-van de Geijn, H. M., Hoogland,<br />

A. E. & Vereijken, P. F. G. (2003). Quantitative aspects<br />

of infection of Sclerotinia sclerotiorum sclerotia by<br />

Coniothyrium minitans timing of application,<br />

concentration and quality of conidial suspension of<br />

the mycoparasite. European Journal of Plant Pathology,<br />

109, 489 502.<br />

Gernandt, D. S. & S<strong>to</strong>ne, J. K. (1999). Phylogenetic<br />

analysis of nuclear ribosomal DNA places the<br />

nema<strong>to</strong>de parasite, Drechmeria coniospora, in<br />

Clavicipitaceae. Mycologia, 91, 993 1000.<br />

Gernandt, D. S., Platt, J. L., S<strong>to</strong>ne, J. K., et al. (2001).<br />

Phylogenetics of Helotiales and Rhytismatales based<br />

on partial small subunit nuclear ribosomal DNA<br />

sequences. Mycologia, 93, 915 933.<br />

Gessner, M. O., Suberkropp, K. & Chauvet, E. (1997).<br />

Decomposition of plant litter by fungi in marine and<br />

freshwater ecosystems. In The Mycota IV:<br />

Environmental and Microbial Relationships, ed. D. T.<br />

Wicklow & B. Söderström. Berlin: Springer-Verlag,<br />

pp. 302 332.<br />

Gessner, R. V. (1995). Genetics and systematics of North<br />

American populations of Morchella. Canadian Journal<br />

of Botany, 73, S967 S972.<br />

Gessner, R. V., Romano, M. A. & Schulz, R. W. (1987).<br />

Allelic variation and segregation in Morchella deliciosa<br />

and M. esculenta. Mycologia, 79, 683 687.<br />

Ghannoum, M. A. & Rice, L. B. (1999). Antifungal agents:<br />

mode of action, mechanisms of resistance, and<br />

correlation of these mechanisms with bacterial<br />

resistance. Clinical Microbiology Reviews, 12, 501 517.<br />

Gibbs, J. N. (1993). The biology of ophios<strong>to</strong>moid fungi<br />

causing sapstain in trees and freshly cut logs. In<br />

Cera<strong>to</strong>cystis and Ophios<strong>to</strong>ma. Taxonomy, Ecology and<br />

Pathogenicity, ed. M. J. Wingfield, K. A. Seifert & J. F.<br />

Webber. St. Paul, USA: APS Press, pp. 153 160.<br />

Giese, H., Hippe-Sanwald, S., Somerville, S. & Weller, J.<br />

(1997). Erysiphe graminis. InThe Mycota VB: Plant<br />

Relationships, ed. G. C. Carroll & P. Tudzynski. Berlin:<br />

Springer-Verlag, pp. 55 78.


738 REFERENCES<br />

Giesy, R. M. & Day, P. R. (1965). The septal pores of<br />

Coprinus lagopus (Fr.) sensu Buller in relation <strong>to</strong> nuclear<br />

migration. American Journal of Botany, 52, 287 293.<br />

Gil, F. & Gay, J. L. (1977). Ultrastructural and physiological<br />

properties of the host interfacial components<br />

of haus<strong>to</strong>ria of Erysiphe pisi in vivo and in vitro.<br />

Physiological Plant Pathology, 10, 1 10.<br />

Gil-ad, N. L., Bar-Nun, N. & Mayer, A. M. (2001). The<br />

possible function of the glucan sheath of Botrytis<br />

cinerea: effects on the distribution of enzyme<br />

activities. FEMS Microbiology Letters, 199, 109 113.<br />

Gilijamse, E., Frinking, H. D. & Jeger, M. J. (1997).<br />

Occurrence and epidemiology of pearl millet downy<br />

mildew, Sclerospora graminicola, in southwest Niger.<br />

International Journal of Pest Management, 43, 279 283.<br />

Gill, M. & Steglich, W. (1987). Pigments of <strong>Fungi</strong><br />

(Macromycetes). Vienna: Springer-Verlag.<br />

Gill, M. & Strauch, R. J. (1984). Constituents of Agaricus<br />

xanthodermus Genevier: the first naturally endogenous<br />

azo compound and <strong>to</strong>xic phenolic metabolites.<br />

Zeitschrift für Naturforschung, 39c, 1027 1029.<br />

Gill, M. & Watling, R. (1986). The relationships of<br />

Pisolithus (Sclerodermataceae) <strong>to</strong> other fleshy fungi<br />

with particular reference <strong>to</strong> the occurrence and<br />

taxonomic significance of hydroxylated pulvinic<br />

acids. Plant Systematics and Evolution, 154, 225 236.<br />

Gilles, T., Ashby, A., Fitt, B. D. L. & Cole, T. (2001).<br />

Development of Pyrenopeziza brassicae apothecia on<br />

agar and oilseed rape debris. Mycological Research, 105,<br />

705 714.<br />

Gillis, A. M. (1993). The magnificent devasta<strong>to</strong>r gets<br />

around. Bioscience, 43, 368 371.<br />

Gimeno, C. J., Ljungdahl, P. O., Styles, C. A. & Fink, G. R.<br />

(1992). Unipolar cell divisions in the yeast S. cerevisiae<br />

lead <strong>to</strong> filamen<strong>to</strong>us growth: regulation by starvation<br />

and RAS. Cell, 68, 1077 1090.<br />

Gindin, G. & Ben Ze’ev, I. S. (1994). Natural occurrence<br />

of and inoculation experiments with Conidiobolus<br />

coronatus and Conidiobolus sp. in glasshouse populations<br />

of Bemisia tabaci. Phy<strong>to</strong>parasitica, 22, 187 208.<br />

Ginman, A. & Young, T. W. K. (1989). Azygospore<br />

morphology in Mucor azygospora and M. bainieri.<br />

Mycological Research, 93, 314 320.<br />

Giollant, M., Guillot, J., Damez, M., et al. (1993).<br />

Characterization of a lectin from Lactarius deterrimus:<br />

research on the possible involvement of the fungal<br />

lectin in recognition between mushroom and spruce<br />

during the early stages of mycorrhizae formation.<br />

Plant Physiology, 101, 513 522.<br />

Giovannetti, G., Roth-Bejerano, N., Zanini, E. & Kagan-<br />

Zur, V. (1994). Truffles and their cultivation.<br />

Horticultural Reviews, 16, 71 107.<br />

Giraud, T., Fortini, D., Levis, C., et al. (1999). Two sibling<br />

species of the Botrytis cinerea complex, transposa and<br />

vacuma, are found in sympatry on numerous host<br />

plants. Phy<strong>to</strong>pathology, 89, 967 973.<br />

Girbardt, M. (1968). The ultrastructure and dynamics of<br />

the moving nucleus. Symposia of the Society for<br />

Experimental Biology, 22, 249 259.<br />

Girbardt, M. (1969). Die Ultrastruktur der Apikalregion<br />

von Pilzhyphen. Pro<strong>to</strong>plasma, 67, 413 441.<br />

Gisi, U. (1983). Biophysical aspects of the development<br />

of Phy<strong>to</strong>phthora. InPhy<strong>to</strong>phthora. Its Biology, Taxonomy,<br />

Ecology, and Pathology, ed. D. C. Erwin, S. Bartnicki-<br />

Garcia & P. H. Tsao. St. Paul: APS Press, pp. 109 119.<br />

Giuma, A. Y., Hackett, A. M. & Cooke, R. C. (1973).<br />

Thermostable nema<strong>to</strong><strong>to</strong>xins produced by germinating<br />

conidia of some endozoic fungi. Transactions of the<br />

British Mycological Society, 60, 49 56.<br />

Glass, N. L. & Lorimer, A. J. (1991). Ascomycete mating<br />

type genes. In More Gene Manipulations in <strong>Fungi</strong>, ed.<br />

J. W. Bennett & L. L. Lasure. San Diego: Academic<br />

Press, pp. 194 216.<br />

Glass, N. L., Vollmer, S. J., Staben, C., et al. (1988). DNAs<br />

of the two mating-type alleles of Neurospora crassa are<br />

highly dissimilar. Science, 241, 570 573.<br />

Glass, N. L., Grötelueschen, J. & Metzenberg, R. L. (1990).<br />

Neurospora crassa A mating-type region. Proceedings of<br />

the National Academy of Sciences of the USA, 87,<br />

4912 4916.<br />

Glass, N. L., Jacobson, D. J. & Shiu, P. K. T. (2000). The<br />

genetics of hyphal fusion and vegetative incompatibility<br />

in filamen<strong>to</strong>us ascomycete fungi. Annual<br />

Review of Genetics, 34, 165 186.<br />

Gleason, F. H. (1976). The physiology of the lower<br />

freshwater fungi. In Recent Advances in Aquatic<br />

Mycology, ed. E. B. Gareth-Jones. London: Elek Science,<br />

pp. 543 572.<br />

Gleason, F. H., Letcher, P. M. & McGee, P. A. (2004). Some<br />

Chytridiomycota in soil recover from drying and high<br />

temperatures. Mycological Research, 108, 583 589.<br />

Glenn, A. E., Bacon, C. W., Price, R. & Hanlin, R. T. (1996).<br />

Molecular phylogeny of Acremonium and its taxonomic<br />

implications. Mycologia, 88, 369 383.<br />

Glockling, S. L. (1998). Isolation of a new species of<br />

rotifer-attacking Olpidium. Mycological Research, 102,<br />

206 208.<br />

Glockling, S. L. & Beakes, G. W. (2000a). Video microscopy<br />

of spore development in Hap<strong>to</strong>glossa heteromorpha,<br />

a new species from cow dung. Mycologia, 92,<br />

747 753.<br />

Glockling, S. L. & Beakes, G. W. (2000b). An ultrastructural<br />

study of sporidium formation during infection<br />

of a rhabditid nema<strong>to</strong>de by large gun cells of


REFERENCES<br />

739<br />

Hap<strong>to</strong>glossa heteromorpha. Journal of Invertebrate<br />

Pathology, 76, 208 215.<br />

Goates, B. J. (1988). His<strong>to</strong>logy of infection of wheat by<br />

Tilletia indica, the Karnal bunt pathogen.<br />

Phy<strong>to</strong>pathology, 78, 1434 1441.<br />

Godoy, G., Steadman, J. R., Dickman, M. B. & Dam, R.<br />

(1990). Use of mutants <strong>to</strong> demonstrate the role of<br />

oxalic acid in pathogenicity of Sclerotinia sclerotiorum<br />

on Phaseolus vulgaris. Physiological and Molecular Plant<br />

Pathology, 37, 179 191.<br />

Gold, R. E. & Mendgen, K. (1991). Rust basidiospore<br />

germlings and disease initiation. In The Fungal Spore<br />

and Disease Initiation in Plants and Animals, ed. G. T.<br />

Cole & H. C. Hoch. New York: Plenum Press,<br />

pp. 67 99.<br />

Gold, S. E., Brogdon, S. M., Mayorga, M. E. & Kronstad,<br />

J. W. (1997). The Ustilago maydis regula<strong>to</strong>ry subunit<br />

of a cAMP-dependent protein kinase is required<br />

for gall formation in maize. Plant Cell, 9, 1585 1594.<br />

Goldstein, B. (1923). Resting spores of Empusa muscae.<br />

Bulletin of the Torrey Botanical Club, 50, 317 327.<br />

Goldstein, S. (1960). Fac<strong>to</strong>rs affecting the growth and<br />

pigmentation of Cladochytrium replicatum. Mycologia,<br />

52, 490 498.<br />

Goldstein, S. (1961). Studies of two polycentric chytrids<br />

in pure culture. American Journal of Botany, 48,<br />

294 298.<br />

Gönczöl, J. & Révay, A. (2003). Treehole fungal<br />

communities: aquatic, aero-aquatic and<br />

dematiaceous hyphomycetes. Fungal Diversity,<br />

12, 19 34.<br />

Gonzalez, D., Carling, D. E., Kuninaga, S., Vilgalys, R. &<br />

Cubeta, M. A. (2001). Ribosomal DNA systematics of<br />

Cera<strong>to</strong>basidium and Thanatephorus with Rhizoc<strong>to</strong>nia<br />

anamorphs. Mycologia, 93, 1138 1150.<br />

Gooday, G. W. (1994). Hormones in mycelial fungi. In<br />

The Mycota I: Growth, Differentiation and Sexuality, ed.<br />

J. G. H. Wessels & F. Meinhardt. Berlin: Springer-<br />

Verlag, pp. 401 12.<br />

Gooday, G. W. (1995). Cell walls. In The Growing Fungus,<br />

ed. N. A. R. Gow & G. M. Gadd. London: Chapman &<br />

Hall, pp. 45 62.<br />

Gooday, G. W. & Adams, D. J. (1992). Sex hormones and<br />

fungi. Advances in Microbial Physiology, 34, 69 145.<br />

Gooday, G. W. & Carlile, M. J. (1997). The discovery of<br />

fungal sex hormones: III. Trisporic acid and its<br />

precursors. Mycologist, 11, 126 130.<br />

Gooday, G. W., Fawcett, P., Green, D. & Shaw, D. (1973).<br />

The formation of sporopollenin in the zygospore<br />

wall of Mucor mucedo: a role for the sexual carotenogenesis<br />

in the Mucorales. Journal of General<br />

Microbiology, 74, 233 239.<br />

Goodwin, S. B. (2002). The barley scald pathogen<br />

Rhynchosporium secalis is closely related <strong>to</strong> the<br />

discomycetes Tapesia and Pyrenopeziza. Mycological<br />

Research, 106, 645 654.<br />

Goodwin, S. B. & Zismann, V. L. (2001). Phylogenetic<br />

analyses of the ITS region of ribosomal DNA reveal<br />

that Sep<strong>to</strong>ria passerinii from barley is closely related<br />

<strong>to</strong> the wheat pathogen Mycosphaerella graminicola.<br />

Mycologia, 93, 934 946.<br />

Goodwin, S. B., Cohen, B. A. & Fry, W. E. (1994a).<br />

Panglobal distribution of a single clonal lineage<br />

of the Irish pota<strong>to</strong> famine fungus. Proceedings of<br />

the National Academy of Sciences of the USA, 91,<br />

11 591 595.<br />

Goodwin, S. B., Cohen, B. A., Deahl, K. L. & Fry, W. E.<br />

(1994b). Migration from northern Mexico as the<br />

probable cause of recent genetic changes in populations<br />

of Phy<strong>to</strong>phthora infestans in the United States<br />

and Canada. Phy<strong>to</strong>pathology, 84, 553 558.<br />

Goodwin, S. B., Sujkowski, L. S., Dyer, A. T., Fry, B. A. &<br />

Fry, W. E. (1995). Direct detection of gene flow and<br />

probable sexual reproduction of Phy<strong>to</strong>phthora infestans<br />

in northern North America. Phy<strong>to</strong>pathology, 85,<br />

473 479.<br />

Goodwin, S. B., Dunkle, L. & Zismann, V. L. (2001).<br />

Phylogenetic analysis of Cercospora and Mycosphaerella<br />

based on the internal transcribed spacer region of<br />

ribosomal DNA. Phy<strong>to</strong>pathology, 91, 648 658.<br />

Goos, R. D. (1987). <strong>Fungi</strong> with a twist. The helicosporous<br />

hyphomycetes. Mycologia, 79, 1 22.<br />

Gordon, C. C. (1966). A re-interpretation of the on<strong>to</strong>geny<br />

of the ascocarp of species of the Erysiphaceae.<br />

American Journal of Botany, 53, 652 662.<br />

Gordon, T. R. (1993). Genetic variation and adaptive<br />

potential in an asexual soil-borne fungus. In The<br />

Fungal Holomorph. Mi<strong>to</strong>tic, Meiotic and Pleomorphic<br />

Speciation in Fungal Systematics, ed. D. R. Reynolds &<br />

J. W. Taylor. Wallingford, UK: CAB International,<br />

pp. 217 224.<br />

Gordon, T. R. & Martyn, R. D. (1997). The evolutionary<br />

biology of Fusarium oxysporum. Annual Review of<br />

Phy<strong>to</strong>pathology, 35, 112 128.<br />

Gottlieb, A. M. & Lichtwardt, R. W. (2001). Molecular<br />

variation within and among species of Harpellales.<br />

Mycologia, 93, 66 81.<br />

Govindan, B., Bowser, R. & Novick, P. (1995). The role of<br />

Myo2, a yeast class V myosin, in vesicular transport.<br />

Journal of Cell Biology, 128, 1055 1069.<br />

Govrin, E. M. & Levine, A. (2000). The hypersensitive<br />

response facilitates plant infection by the necrotrophic<br />

pathogen Botrytis cinerea. Current Biology, 10,<br />

751 757.


740 REFERENCES<br />

Gow, N. A. R. (1995). Yeast hyphal dimorphism. In The<br />

Growing Fungus, ed. N. A. R. Gow & G. M. Gadd.<br />

London: Chapman & Hall, pp. 403 422.<br />

Gow, N. A. R. (2002). Candida albicans switches mates.<br />

Molecular Cell, 10, 217 218.<br />

Gow, N. A. R. & Gooday, G. W. (1987). Effects of<br />

antheridiol on growth, branching and electrical<br />

currents of hyphae of Achlya ambisexualis. Journal of<br />

General Microbiology, 133, 3531 3535.<br />

Gow, N. A. R., Bates, S., Brown, A. J., et al. (1999). Candida<br />

cell wall mannosylation: importance in host fungus<br />

interaction and potential as a target for the<br />

development of antifungal drugs. Biochemical Society<br />

Transactions, 27, 512 516.<br />

Graca, M. A. S., Maltby, L. & Calow, P. (1993). Importance<br />

of fungi in the diet of Gammarus pulex and Asellus<br />

aquaticus. II. Effects on growth, reproduction and<br />

physiology. Oecologia, 96, 304 309.<br />

Gradmann, D., Hansen, U.-P., Long, W. S., Slayman, C. L.<br />

& Warncke, J. (1978). Current voltage relationships<br />

for the plasma membrane and its principal electrogenic<br />

pump in Neurospora crassa. I. Steady-state<br />

conditions. Journal of Membrane Biology, 39, 333 367.<br />

Graham, T. R. & Emr, S. D. (1991). Compartmental<br />

organization of Golgi-specific protein modification<br />

and vacuolar protein sorting events defined in a<br />

yeast sec18 (NSF) mutant. Journal of Cell Biology, 114,<br />

207 218.<br />

Grainger, J. (1962). Vegetative and fructifying growth<br />

in Phallus impudicus. Transactions of the British<br />

Mycological Society, 45, 147 155.<br />

Gräser, Y., el Fari, M., Vilgalys, R., et al. (1999).<br />

Phylogeny and taxonomy of the family<br />

Arthrodermataceae (derma<strong>to</strong>phytes) using sequence<br />

analysis of the ribosomal ITS region. Medical<br />

Mycology, 37, 105 114.<br />

Gray, N. F. (1987). Nema<strong>to</strong>phagous fungi with particular<br />

reference <strong>to</strong> their ecology. Biological Reviews, 62,<br />

245 304.<br />

Green, F. & Clausen, C. A. (1999). Production of<br />

polygalacturonase and increase of longitudinal gas<br />

permeability in Southern Pine by brown-rot and<br />

white-rot fungi. Holzforschung, 53, 563 568.<br />

Green, F. & Clausen, C. A. (2003). Copper <strong>to</strong>lerance of<br />

brown-rot fungi: time course of oxalic acid production.<br />

International Biodeterioration and Biodegradation,<br />

51, 145 149.<br />

Green, F., Larsen, M. J., Winandy, J. E. & Highley, T. L.<br />

(1991). Role of oxalic acid in incipient brown-rot<br />

decay. Material und Organismen, 26, 191 213.<br />

Green, G. J. (1971). Hybridization between Puccinia<br />

graminis tritici and Puccinia graminis secalis and its<br />

evolutionary implications. Canadian Journal of Botany,<br />

49, 2089 2095.<br />

Green, J. R., Carver, T. L. W. & Gurr, S. J. (2002). The<br />

formation and function of infection and feeding<br />

structures. In The Powdery Mildews. A Comprehensive<br />

Treatise, ed. R. R. Bélanger, W. R. Bushnell,<br />

A. J. Dik & T. L. W. Carver. St. Paul, USA: APS Press,<br />

pp. 66 82.<br />

Greenhalgh, G. N. & Bevan, R. J. (1978). Response of<br />

Rhytisma acerinum <strong>to</strong> air pollution. Transactions of the<br />

British Mycological Society, 71, 491 523.<br />

Greenhalgh, G. N. & Evans, L. V. (1967). The structure of<br />

the ascus apex in Hypoxylon fragiforme with reference<br />

<strong>to</strong> ascospore release in this and related species.<br />

Transactions of the British Mycological Society, 50,<br />

183 188.<br />

Gregory, P. H. (1949). The operation of the puff-ball<br />

mechanism of Lycoperdon perlatum by raindrops<br />

shown by ultra-high-speed Schlieren cinema<strong>to</strong>graphy.<br />

Transactions of the British Mycological Society, 32,<br />

11 15.<br />

Gregory, P. H. (1957). Electrostatic charges on spores of<br />

fungi in air. Nature, 180, 330.<br />

Gregory, P. H. (1966). The fungus spore: what it is and<br />

what it does. In The Fungal Spore, ed. M. F. Madelin.<br />

London: Butterworths, pp. 1 14.<br />

Gregory, P. H. (1973). Microbiology of the Atmosphere, 2nd<br />

edn. Aylesbury, Bucks: Leonard Hill.<br />

Gregory, P. H. & Sreeramulu, T. (1958). Air spora of an<br />

estuary. Transactions of the British Mycological Society,<br />

41, 145 156.<br />

Greif, M. D. & Currah, R. S. (2003). A functional<br />

interpretation of the role of the reticuloperidium in<br />

whole-ascoma dispersal by arthropods. Mycological<br />

Research, 107, 77 81.<br />

Grente, J. & Sauret, S. (1969). L’hypovirulence exclusive,<br />

phenomène original en pathologie végétale. Comptes<br />

Rendus Hebdomadaires des Séances de l’Académie des<br />

Sciences, D 286, 2347 2350.<br />

Grenville, D. J., Peterson, R. L. & Piché, Y. (1985). The<br />

development, structure, and his<strong>to</strong>chemistry of<br />

sclerotia of ec<strong>to</strong>mycorrhizal fungi. I. Pisolithus<br />

tinc<strong>to</strong>rius. Canadian Journal of Botany, 63, 1402 1411.<br />

Gries, C. (1996). Lichens as indica<strong>to</strong>rs of air pollution.<br />

In Lichen Biology, ed. T. H. Nash. Cambridge:<br />

Cambridge University Press, pp. 240 254.<br />

Griffin, D. H. (1994). Fungal Physiology, 2nd edn. New<br />

York: Wiley-Liss.<br />

Griffith, G. W., Nicholson, J., Nenninger, A., Birch, R. N.<br />

& Hedger, J. N. (2003). Witches’ brooms and frosty<br />

pods: two major pathogens of cacao. New Zealand<br />

Journal of Botany, 41, 423 435.


REFERENCES<br />

741<br />

Griffith, J. M., Davis, A. J. & Grant, B. R. (1992). Target<br />

sites of fungicides <strong>to</strong> control Oomycetes. In Target<br />

Sites of <strong>Fungi</strong>cide Action, ed. W. Köller. Boca Ra<strong>to</strong>n: CRC<br />

Press, pp. 69 100.<br />

Griffiths, A. J. F. (1992). Fungal senescence. Annual<br />

Review of Genetics, 26, 351 372.<br />

Griffiths, D. A. (1974). The origin, structure and function<br />

of chlamydospores in fungi. Nova Hedwigia, 25,<br />

503 547.<br />

Grime, J. P., Mackey, J. M. L., Hillier, S. H. & Read, S. J.<br />

(1987). Floristic diversity in a model system using<br />

experimental microcosms. Nature, 328, 420 422.<br />

Groff, J. M., Mughannam, A., McDowell, T. S., et al.<br />

(1991). An epizootic of cutaneous zygomycosis in<br />

cultured dwarf African clawed frogs (Hymenochirus<br />

curtipes) due <strong>to</strong> Basidiobolus ranarum. Journal of Medical<br />

and Veterinary Mycology, 29, 215 223.<br />

Grogan, R. G., Zink, F. W., Hewitt, W. B. & Kimble, K. A.<br />

(1958). The association of Olpidium with the big-vein<br />

disease of lettuce. Phy<strong>to</strong>pathology, 48, 292 297.<br />

Gross, G. (1987). Zu europäischen Sippen der Gattung<br />

Tuber. InAtlas der Pilze des Saarlandes 2, ed. H. Derbsch<br />

& J. A. Schmitt. Saarbrücken: Verlag der Derlattinia,<br />

pp. 79 100.<br />

Grove, S. N., Bracker, C. E. & Morré, D. J. (1968).<br />

Cy<strong>to</strong>membrane differentiation in the endoplasmic<br />

reticulum Golgi apparatus vesicle complex.<br />

Science, 161, 171 173.<br />

Grove, W. B. (1913). The British Rust <strong>Fungi</strong> (Uredinales).<br />

Their Biology and Classification. Cambridge: Cambridge<br />

University Press.<br />

Grove, W. B. (1934). A systematic account and arrangement<br />

of the Pilobolidae. In Buller, A. H. R.: Researches<br />

on <strong>Fungi</strong>, 6, pp. 190 224. London: Longmans, Green<br />

& Co.<br />

Grove, W. B. (1935). British Stem- and Leaf-<strong>Fungi</strong><br />

(Coelomycetes) I. Cambridge: Cambridge University<br />

Press.<br />

Grsic, S., Kirchheim, B., Pieper, K., et al. (1999).<br />

Induction of auxin biosynthetic enzymes by jasmonic<br />

acid and in clubroot diseased Chinese cabbage<br />

plants. Physiologia Plantarum, 105, 521 531.<br />

Grube, M. & Winka, K. (2002). Progress in understanding<br />

the evolution and classification of lichenized<br />

ascomycetes. Mycologist, 16, 67 76.<br />

Grubisha, L. C., Trappe, J. M., Molina, R. & Spatafora,<br />

J. W. (2001). Biology of the ec<strong>to</strong>mycorrhizal genus<br />

Rhizopogon. V. Phylogenetic relationships in the<br />

Boletales inferred from LSU rDNA sequences.<br />

Mycologia, 93, 82 89.<br />

Grylls, B. T. & Seifert, K. A. (1993). A synoptic key <strong>to</strong><br />

species of Ophios<strong>to</strong>ma, Cera<strong>to</strong>cystis and Cera<strong>to</strong>cystiopsis.<br />

In Cera<strong>to</strong>cystis and Ophios<strong>to</strong>ma. Taxonomy, Ecology and<br />

Pathogenicity, ed. M. J. Wingfield, K. A. Seifert & J. F.<br />

Webber. St. Paul, USA: APS Press, pp. 261 268.<br />

Gu, Y. H. & Ko, W. H. (1998). Occurrence of a parasexual<br />

cycle following the transfer of isolated nuclei in<strong>to</strong><br />

pro<strong>to</strong>plasts of Phy<strong>to</strong>phthora parasitica. Current Genetics,<br />

34, 120 123.<br />

Guarro, J. & Figueras, M. J. (1989). Morphological<br />

studies in Chae<strong>to</strong>mium: the peridium, the ostiolar<br />

apparatus and some considerations about the<br />

centrum. Cryp<strong>to</strong>gamic Botany, 1, 97 140.<br />

Guarro, J. & von Arx, J. A. (1987). The ascomycete genus<br />

Sordaria. Persoonia, 13, 301 313.<br />

Gubler, F. & Hardham, A. R. (1988). Secretion of<br />

adhesive material during encystment of Phy<strong>to</strong>phthora<br />

cinnamomi zoospores, characterized by immunogold<br />

labeling with monoclonal antibodies <strong>to</strong> components<br />

of peripheral vesicles. Journal of Cell Science, 90,<br />

225 235.<br />

Gubler, F. & Hardham, A. R. (1990). Protein s<strong>to</strong>rage in<br />

large peripheral vesicles in Phy<strong>to</strong>phthora zoospores<br />

and its breakdown after cyst germination.<br />

Experimental Mycology, 14, 393 404.<br />

Gucia, M. & Falandysz, J. (2003). Total mercury content<br />

in parasol mushroom Macrolepiota procera from<br />

various sites in Poland. Journal de Physique IV, 107,<br />

581 584.<br />

Guého, E. & Meyer, S. A. (1989). A reevaluation of the<br />

genus Malassezia by means of genome comparison.<br />

An<strong>to</strong>nie van Leeuwenhoek, 55, 245 251.<br />

Guého, E., Midgley, G. & Guillot, J. (1996). The genus<br />

Malassezia with description of four new species.<br />

An<strong>to</strong>nie van Leeuwenhoek, 69, 337 355.<br />

Guého, E., Leclerc, M. C., de Hoog, G. S. & Dupont, B.<br />

(1997). Molecular taxonomy and epidemiology of<br />

Blas<strong>to</strong>myces and His<strong>to</strong>plasma species. Mycoses, 40,<br />

69 81.<br />

Guerin-Languette, A., Matsushita, N., Lapeyrie, F.,<br />

Shindo, K. & Suzuki, K. (2005). Successful inoculation<br />

of mature pine with Tricholoma matsutake. Mycorrhiza,<br />

15, 301 305.<br />

Guerrero, A., Jain, N., Goldman, D. L. & Fries, B. C.<br />

(2006). Phenotypic switching in Cryp<strong>to</strong>coccus neoformans.<br />

Microbiology, 152, 3 9.<br />

Gugnani, H. C. (1992). En<strong>to</strong>mophthoromycosis due <strong>to</strong><br />

Conidiobolus. European Journal of Epidemiology, 8,<br />

391 396.<br />

Gugnani, H. C. (1999). A review of zygomycosis due <strong>to</strong><br />

Basidiobolus ranarum. European Journal of Epidemiology,<br />

15, 923 929.<br />

Guillot, J. & Bond, R. (1999). Malassezia pachydermatis: a<br />

review. Medical Mycology, 37, 295 306.


742 REFERENCES<br />

Guillot, J. & Guého, E. (1995). The diversity of Malassezia<br />

yeasts confirmed by rRNA sequence and nuclear DNA<br />

comparisons. An<strong>to</strong>nie van Leeuwenhoek, 67, 297 314.<br />

Gulati, A., Gulati, A., Ravindranath, S. D. & Gupta, A. K.<br />

(1999). Variation in chemical composition and<br />

quality of tea (Camellia sinensis) with increasing<br />

blister blight (Exobasidium vexans) severity. Mycological<br />

Research, 103, 1380 1384.<br />

Gulis, V. (2001). Are there any substrate preferences in<br />

aquatic hyphomycetes? Mycological Research, 105,<br />

1088 1093.<br />

Gull, K. (1978). Form and function of septa in<br />

filamen<strong>to</strong>us fungi. In The Filamen<strong>to</strong>us <strong>Fungi</strong>, 3.<br />

Developmental Mycology, ed. J. E. Smith & D. R. Berry.<br />

London: Edward Arnold, pp. 78 93.<br />

Guo, L.-Y. & Michailides, T. J. (1998). Fac<strong>to</strong>rs affecting<br />

the rate and mode of germination of Mucor piriformis<br />

zygospores. Mycological Research, 102, 815 819.<br />

Gupta, A. K., Einarson, T. R., Summerbell, R. C. & Shear,<br />

N. H. (1998). An overview of <strong>to</strong>pical antifungal<br />

therapy in derma<strong>to</strong>mycoses. A North American<br />

perspective. Drugs, 55, 645 674.<br />

Gupta, A. K., Bluhm, R. & Summerbell, R. (2002).<br />

Pityriasis versicolor. Journal of the European Academy of<br />

Derma<strong>to</strong>logy and Venereology, 16, 19 33.<br />

Guttenberger, M. (2000). Arbuscules of vesicular<br />

arbuscular fungi inhabit an acidic compartment<br />

within plant roots. Planta, 211, 299 304.<br />

Gwynne, D. I. (1992). Foreign proteins. In Applied<br />

Molecular Genetics of Filamen<strong>to</strong>us <strong>Fungi</strong>, ed. J. R.<br />

Kinghorn & G. Turner. Glasgow: Blackie, pp. 132 51.<br />

Haase, G., Sonntag, L., Melzer-Krick, B. & de Hoog, G. S.<br />

(1999). Phylogenetic inference by SSU-gene analysis<br />

of members of the Herpotrichiellaceae with special<br />

reference <strong>to</strong> human pathogenic species. Studies in<br />

Mycology, 43, 80 97.<br />

Haber, J. E. (1998). Mating-type gene switching in<br />

Saccharomyces cerevisiae. Annual Review of Genetics, 32,<br />

561 599.<br />

Hadley, G. (1968). Development of stromata in Claviceps<br />

purpurea. Transactions of the British Mycological Society,<br />

51, 763 769.<br />

Hadley, G. (1970). Non-specificity of symbiotic infection<br />

in orchid mycorrhiza. New Phy<strong>to</strong>logist, 69, 1015 1023.<br />

Hadley, G. & Williamson, B. (1971). Analysis of the postinfection<br />

growth stimulus in orchid mycorrhiza. New<br />

Phy<strong>to</strong>logist, 70, 445 455.<br />

Hagan, I. M. (1998). The fission yeast microtubule<br />

cy<strong>to</strong>skele<strong>to</strong>n. Journal of Cell Science, 111, 1603 1612.<br />

Hagedorn, G. & Scholler, M. (1999). A reevaluation of<br />

preda<strong>to</strong>ry orbiliaceous fungi. I. Phylogenetic analysis<br />

using rDNA sequence data. Sydowia, 51, 27 48.<br />

Hahn, M. (2000). The rust fungi. In Fungal Pathology, ed.<br />

J. W. Kronstad. Dordrecht: Kluwer, pp. 267 306.<br />

Hajek, A. E., Wraight, S. P. & Vandenberg, J. D. (2001).<br />

Control of arthropods using pathogenic fungi. In Bio-<br />

Exploitation of Filamen<strong>to</strong>us <strong>Fungi</strong>, ed. S. B. Pointing &<br />

K. D. Hyde. Hong Kong: Fungal Diversity Press,<br />

pp. 309 347.<br />

Hale, M. E. (1983). The Biology of Lichens, 3rd edn.<br />

Baltimore: Edward Arnold.<br />

Hall, I. R. (1984). Taxonomy of VA mycorrhizal fungi. In<br />

VA Mycorrhiza, ed. C. L. Powell & D. J. Bagyaraj. Boca<br />

Ra<strong>to</strong>n: CRC Press, pp. 57 94.<br />

Hall, I. R., Brown, G. & Byars, J. (1994). The Black Truffle,<br />

its His<strong>to</strong>ry, Uses and Cultivation. Lincoln, New Zealand:<br />

Crop and Food Research.<br />

Hall, I. R., Yun, W. & Amicucci, A. (2003). Cultivation of<br />

edible ec<strong>to</strong>mycorrhizal mushrooms. Trends in<br />

Biotechnology, 21, 433 438.<br />

Hallett, I. C. & Dick, M. W. (1986). Fine structure of<br />

zoospore cyst ornamentation in the Saprolegniaceae<br />

and Pythiaceae. Transactions of the British Mycological<br />

Society, 86, 457 463.<br />

Hamer, J. E., Howard, R. J., Chumley, F. G. & Valent, B.<br />

(1988). A mechanism for surface attachment in spores<br />

of a plant pathogenic fungus. Science, 239, 288 290.<br />

Hammett, K. R. W. & Manners, J. G. (1971). Conidium<br />

liberation in Erysiphe graminis. I. Visual and statistical<br />

analysis of spore trap records. Transactions of the<br />

British Mycological Society, 56, 387 401.<br />

Hammett, K. R. W. & Manners, J. G. (1974). Conidium<br />

liberation in Erysiphe graminis. III. Wind tunnel<br />

studies. Transactions of the British Mycological Society,<br />

62, 267 282.<br />

Hammill, T. M. (1981). Mucoralean sporangiosporogenesis.<br />

In The Fungal Spore: Morphogenetic Controls, ed. G.<br />

Turian & H. R. Hohl. New York: Academic Press,<br />

pp. 173 194.<br />

Hammill, T. M. & Secor, D. L. (1983). The number of<br />

nuclei in sporangiospores of Mucor mucedo. Mycologia,<br />

75, 648 655.<br />

Hammond, J. B. W. (1985). The biochemistry of Agaricus<br />

fructification. In Developmental Biology of Higher <strong>Fungi</strong>,<br />

ed. D. Moore, L. A. Cassel<strong>to</strong>n, D. A. Wood & J. C.<br />

Frankland. Cambridge: Cambridge University Press,<br />

pp. 339 401.<br />

Hammond, J. B. W. & Nichols, R. (1979). Carbohydrate<br />

metabolism in Agaricus bisporus: changes in nonstructural<br />

carbohydrates during periodic fruiting<br />

(flushing). New Phy<strong>to</strong>logist, 83, 723 730.<br />

Hammond, J. R. (1995). Genetically-modified brewing<br />

yeasts for the 21st century. Progress <strong>to</strong> date. Yeast, 11,<br />

1613 1627.


REFERENCES<br />

743<br />

Hampson, M. C. (1988). Control of pota<strong>to</strong> wart disease<br />

through the application of chemical soil treatments:<br />

an his<strong>to</strong>rical review of earlier studies (1909 1928).<br />

EPPO Bulletin, 18, 153 161.<br />

Hampson, M. C. & Coombes, J. W. (1991). Use of<br />

crabshell meal <strong>to</strong> control pota<strong>to</strong> wart in<br />

Newfoundland. Canadian Journal of Plant Pathology, 13,<br />

97 105.<br />

Hampson, M. C., Yang, A. F. & Bal, A. K. (1994).<br />

Ultrastructure of Synchytrium endobioticum and<br />

enhancement of germination using snails. Mycologia,<br />

86, 733 740.<br />

Hankin, L. & Puhalla, J. E. (1971). Nature of a fac<strong>to</strong>r<br />

causing interstrain lethality in Ustilago maydis.<br />

Phy<strong>to</strong>pathology, 61, 50 53.<br />

Hanlin, R. T. (1971). Morphology of Nectria haema<strong>to</strong>cocca.<br />

American Journal of Botany, 58, 105 116.<br />

Hansen, E. M. & Goheem, E. M. (2000). Phellinus weirii<br />

and other native root pathogens as determinants of<br />

forest structure and process in Western North<br />

America. Annual Review of Phy<strong>to</strong>pathology, 38,<br />

515 539.<br />

Hansen, K., Laessøe, T. & Pfister, D. H. (2001).<br />

Phylogenetics of the Pezizaceae, with an emphasis<br />

on Peziza. Mycologia, 93, 958 990.<br />

Hansen, K., LoBuglio, K. F. & Pfister, D. H. (2005).<br />

Evolutionary relationships of the cup-fungus genus<br />

Peziza and Pezizaceae inferred from multiple nuclear<br />

genes: RPB2, b-tubulin, and LSU rDNA. Molecular<br />

Phylogenetics and Evolution, 36, 1 23.<br />

Hansen, L. (1958). On the ana<strong>to</strong>my of the Danish<br />

species of Ganoderma. Botanisk Tidskrift, 54, 333 352.<br />

Hansen, L. & Knudsen, H., eds. (2000). Nordic<br />

Macromycetes 1. Ascomycetes. Copenhagen:<br />

Nordsvamp.<br />

Hanssen, H.-P. (1993). Volatile metabolites produced by<br />

species of Ophios<strong>to</strong>ma and Cera<strong>to</strong>cystis. In Cera<strong>to</strong>cystis<br />

and Ophios<strong>to</strong>ma. Taxonomy, Ecology and Pathogenicity,<br />

ed. M. J. Wingfield, K. A. Seifert & J. F. Webber. St.<br />

Paul, USA: APS Press, pp. 117 125.<br />

Hantula, J., Kasanen, R., Kaitera, J. & Moricca, S. (2002).<br />

Analyses of genetic variation suggest that pine rusts<br />

Cronartium flaccidum and Peridermium pini belong <strong>to</strong><br />

the same species. Mycological Research, 106, 203 209.<br />

Harder, D. E. & Chong, J. (1984). Structure and<br />

physiology of haus<strong>to</strong>ria. In The Cereal Rusts 1: Origins,<br />

Specificity, Structure, and Physiology, ed. W. R. Bushnell<br />

& A. P. Roelfs. Orlando: Academic Press, pp. 431 476.<br />

Hardham, A. R. (1995). Polarity of vesicle distribution in<br />

oomycete zoospores: development of polarity and<br />

importance for infection. Canadian Journal of Botany,<br />

73, S400 S407.<br />

Hardham, A. R. & Gubler, F. (1990). Polarity of attachment<br />

of zoospores of a root pathogen and prealignment<br />

of the emerging germ tube. Cell Biology<br />

International Reports, 14, 947 956.<br />

Hardham, A. R. & Hyde, G. J. (1997). Asexual sporulation<br />

in the oomycetes. Advances in Botanical Research, 24,<br />

353 398.<br />

Hardham, A. R., Gubler, F. & Duniec, J. (1991).<br />

Ultrastructural and immunological studies of zoospores<br />

of Phy<strong>to</strong>phthora. InPhy<strong>to</strong>phthora, ed. J. A. Lucas,<br />

R. C. Shat<strong>to</strong>ck, D. S. Shaw & L. R. Cooke. Cambridge:<br />

Cambridge University Press, pp. 50 69.<br />

Harold, F. M. (1994). Ionic and electric dimensions of<br />

hyphal growth. In The Mycota I: Growth, Differentiation<br />

and Sexuality, ed. J. G. H. Wessels & F. Meinhardt.<br />

Berlin: Springer-Verlag, pp. 489 509.<br />

Harring<strong>to</strong>n, F. A., Pfister, D. H., Potter, D. & Donoghue,<br />

M. J. (1999). Phylogenetic studies within the<br />

Pezizales. I. 18S rDNA sequence data and classification.<br />

Mycologia, 91, 41 50.<br />

Harring<strong>to</strong>n, T. C. & McNew, D. L. (1997). Self-fertility<br />

and uni-directional mating-type switching in<br />

Cera<strong>to</strong>cystis coerulescens, a filamen<strong>to</strong>us ascomycete.<br />

Current Genetics, 32, 52 59.<br />

Harring<strong>to</strong>n, T. C. & McNew, D. L. (2003). Phylogenetic<br />

analysis places the Phialophora-like anamorph<br />

genus Cadophora in the Helotiales. Mycotaxon, 87,<br />

141 151.<br />

Harring<strong>to</strong>n, T. C. & Wingfield, M. J. (1998). The<br />

Cera<strong>to</strong>cystis species on conifers. Canadian Journal of<br />

Botany, 76, 1446 1457.<br />

Harris, J. S. & Taber, W. A. (1973). Ultrastructure and<br />

morphogenesis of the synnema of Cera<strong>to</strong>cystis ulmi.<br />

Canadian Journal of Botany, 51, 1565 1571.<br />

Harrison, J. G., Searle, R. J. & Williams, N. A. (1997).<br />

Powdery scab of pota<strong>to</strong> a review. Plant Pathology,<br />

46, 1 25.<br />

Harrison, J. S. (1993). Food and fodder yeasts. In The<br />

Yeasts 5: Yeast Technology, ed. A. H. Rose & J. S.<br />

Harrison. London: Academic Press, pp. 399 433.<br />

Harrison, T. S. & Levitz, S. M. (1996). Infections due <strong>to</strong><br />

the dimorphic fungi. In The Mycota VI: Human and<br />

Animal Relationships, ed. D. H. Howard & J. D. Miller.<br />

Berlin: Springer-Verlag, pp. 125 146.<br />

Harrold, C. E. (1950). Studies in the genus Eremascus.<br />

I. The re-discovery of Eremascus albus Eidam and<br />

some new observations concerning its life<br />

his<strong>to</strong>ry and cy<strong>to</strong>logy. Annals of Botany N.S., 14,<br />

127 148.<br />

Harrower, K. M. (1976). The micropycnidiospores of<br />

Lep<strong>to</strong>sphaeria nodorum. Transactions of the British<br />

Mycological Society, 76, 335 336.


744 REFERENCES<br />

Hartley, M. J. & Williams, P. G. (1971). Genotypic<br />

variation within a phenotype as a possible basis for<br />

somatic hybridization in rust fungi. Canadian Journal<br />

of Botany, 49, 1085 1087.<br />

Hartmeier, W. & Reiss, M. (2002). Production of beer<br />

and wine. In The Mycota X: Industrial Applications,<br />

ed. H. D. Osiewacz. Berlin: Springer-Verlag,<br />

pp. 49 65.<br />

Hashmi, M. H., Morgan-Jones, G. & Kendrick, B. (1972).<br />

Conidium on<strong>to</strong>geny in hyphomycetes. Monilia state<br />

of Neurospora si<strong>to</strong>phila and Sclerotinia laxa. Canadian<br />

Journal of Botany, 500, 1461 1463.<br />

Hashmi, M. H., Morgan-Jones, G. & Kendrick, B. (1973).<br />

Conidium on<strong>to</strong>geny in hyphomycetes. The blas<strong>to</strong>conidia<br />

of Cladosporium herbarum and Torula herbarum.<br />

Canadian Journal of Botany, 51, 1089 1091.<br />

Hause, B. & Fester, T. (2005). Molecular and cell biology<br />

of arbuscular mycorrhizal symbiosis. Planta, 221,<br />

184 196.<br />

Hausner, G., Reid, J. & Klassen, G. R. (1993). On the<br />

phylogeny of Ophios<strong>to</strong>ma, Cera<strong>to</strong>cystis s.s., and<br />

Microascus, and relationships within Ophios<strong>to</strong>ma<br />

based on partial ribosomal DNA sequences. Canadian<br />

Journal of Botany, 71, 1249 1265.<br />

Hausner, G., Belkhiri, A. & Klassen, G. R. (2000).<br />

Phylogenetic analysis of the small subunit ribosomal<br />

RNA gene of the hyphochytrid Rhizidiomyces apophysatus.<br />

Canadian Journal of Botany, 78, 124 128.<br />

Hawes, C. R. & Beckett, A. (1977). Conidium on<strong>to</strong>geny in<br />

Thielaviopsis basicola. Transactions of the British<br />

Mycological Society, 68, 304 307.<br />

Hawker, L. E. & Beckett, A. (1971). Fine structure and<br />

development of the zygospore of Rhizopus sexualis<br />

(Smith) Callen. Philosophical Transactions of the Royal<br />

Society, B 263, 71 100.<br />

Hawker, L. E., Thomas, B. & Beckett, A. (1970). An<br />

electron microscope study of structure and germination<br />

of conidia of Cunninghamella elegans. Journal of<br />

General Microbiology, 60, 181 189.<br />

Hawksworth, D. L. (2001). The magnitude of fungal<br />

diversity: the 1.5 million species estimate revisited.<br />

Mycological Research, 105, 1422 1432.<br />

Hawksworth, D. L. & Rose, F. (1970). Qualitative scale<br />

for estimating sulphur dioxide air pollution in<br />

England and Wales using epiphytic lichens. Nature,<br />

227, 145 148.<br />

Hayashi, K., Schoonbek, H.-J. & de Waard, M. A. (2002).<br />

Bcmfs1, a novel major facilita<strong>to</strong>r superfamily transporter<br />

from Botrytis cinerea, provides <strong>to</strong>lerance<br />

<strong>to</strong>wards the natural <strong>to</strong>xic compounds camp<strong>to</strong>thecin<br />

and cercosporin and <strong>to</strong>wards fungicides. Applied and<br />

Environmental Microbiology, 68, 4996 5004.<br />

Hayes,W. A., Randle, P. E. & Last, F. T. (1969). The nature<br />

of the microbial stimulus affecting sporophore<br />

production in Agaricus bisporus (Lange) Singer. Annals<br />

of Applied Biology, 64, 177 187.<br />

Hayles, J. & Nurse, P. (2001). A journey in<strong>to</strong> space.<br />

Nature Reviews in Molecular Cell Biology, 2, 647 656.<br />

Heath, I. B. (1987). Preservation of a labile cortical array<br />

of actin filaments in growing hyphal tips of the<br />

fungus Saprolegnia ferax. European Journal of Cell<br />

Biology, 44, 10 16.<br />

Heath, I. B. (1994). The cy<strong>to</strong>skele<strong>to</strong>n in hyphal growth,<br />

organelle movements, and mi<strong>to</strong>sis. In The Mycota I:<br />

Growth, Differentiation and Sexuality, ed. J. G. H. Wessels<br />

& F. Meinhardt. Berlin: Springer-Verlag, pp. 43 65.<br />

Heath, I. B. (1995a). The cy<strong>to</strong>skele<strong>to</strong>n. In The Growing<br />

Fungus, ed. N. A. R. Gow & G. M. Gadd. London:<br />

Chapman & Hall, pp. 99 134.<br />

Heath, I. B. (1995b). Integration and regulation of<br />

hyphal tip growth. Canadian Journal of Botany, 73,<br />

S131 S139.<br />

Heath, I. B. (2001). Bridging the divide: cy<strong>to</strong>skele<strong>to</strong>n -<br />

plasma membrane cell wall interactions in growth<br />

and development. In The Mycota VIII: Biology of the<br />

Fungal Cell, ed. R. J. Howard & N. A. R. Gow. Berlin:<br />

Springer-Verlag, pp. 201 23.<br />

Heath, I. B. & Harold, R. L. (1992). Actin has multiple<br />

roles in the formation and architecture of zoospores<br />

of the oomycetes, Saprolegnia ferax and Achlya<br />

bisexualis. Journal of Cell Science, 102, 611 627.<br />

Heath, I. B. & Steinberg, G. (1999). Mechanisms of<br />

hyphal tip growth: tube dwelling amoebae revisited.<br />

Fungal Genetics and Biology, 28, 79 93.<br />

Heath, I. B., Greenwood, A. D. & Griffiths, H. B. (1970).<br />

The origin of flimmer in Saprolegnia, Dictyuchus,<br />

Synura and Cryp<strong>to</strong>monas. Journal of Cell Science, 7,<br />

445 461.<br />

Heath, I. B., Bauchop, T. & Skipp, R. K. (1983).<br />

Assignment of the rumen anaerobe Neocallimastix<br />

frontalis <strong>to</strong> the Spizellomycetales (Chytridiomycetes)<br />

on the basis of its polyflagellate zoospore<br />

ultrastructure. Canadian Journal of Botany, 61,<br />

295 307.<br />

Heath, M. C. (1982). Host defense mechanisms against<br />

infection by rust fungi. In The Rust <strong>Fungi</strong>, ed. K. J.<br />

Scott & A. K. Chakravorty. London: Academic Press,<br />

pp. 223 245.<br />

Heath, M. C. (2000). Hypersensitive response-related<br />

death. Plant Molecular Biology, 44, 321 334.<br />

Heath, M. C. (2002). Cellular interactions between<br />

biotrophic fungal pathogens and host or nonhost<br />

plants. Canadian Journal of Plant Pathology, 24,<br />

259 264.


REFERENCES<br />

745<br />

Heath, M. C. & Skalamera, D. (1997). Cellular interactions<br />

between plants and biotrophic fungal parasites.<br />

Advances in Botanical Research, 24, 195 225.<br />

Heath, M. C., Howard, R. J., Valent, B. & Chumley, F. G.<br />

(1992). Ultrastructural interactions of one strain of<br />

Magnaporthe grisea with goosegrass and weeping<br />

lovegrass. Canadian Journal of Botany, 70, 779 787.<br />

Hedge, Y. & Kolattukudy, P. E. (1997). Cuticular waxes<br />

relieve self-inhibition of germination and appressorium<br />

formation by the conidia of Magnaporthe grisea.<br />

Physiological and Molecular Plant Pathology, 51, 75 84.<br />

Hedger, J. N., Lewis, P. & Gitay, H. (1993). Litter-trapping<br />

by fungi in moist tropical forest. In Aspects of Tropical<br />

Mycology, ed. S. Isaac, R. Watling, A. J. S. Whalley &<br />

J. C. Frankland. Cambridge: Cambridge University<br />

Press, pp. 15 35.<br />

Hegarty, B. (1991). Fac<strong>to</strong>rs affecting the fruiting of the<br />

dry rot fungus Serpula lacrymans. In Serpula lacrymans.<br />

Fundamental Biology and Control Strategies, ed. D. H.<br />

Jennings & A. F. Bravery. Chichester: John Wiley &<br />

Sons, pp. 39 53.<br />

Hegarty, B., Steinfurth, A., Liese, W. & Schmidt, O.<br />

(1987). Comparative investigations on wood decay<br />

and cellulolytic and xylanolytic activity of some<br />

basidiomycete fungi. Holzforschung, 41, 265 269.<br />

Hegedus, D. D. & Rimmer, S. R. (2005). Sclerotinia<br />

sclerotiorum: when ‘<strong>to</strong> be or not <strong>to</strong> be’ a pathogen?<br />

FEMS Microbiology Letters, 251, 177 184.<br />

Heim, P. (1960). Évolution du Spongospora parasite des<br />

racines du Cresson. Revue de Mycologie, 25, 3 12.<br />

Heim, R. & Wasson, R. G. (1958). Les Champignons<br />

Hallucinogènes du Mexique. Paris: Musée Nationale<br />

d’His<strong>to</strong>ire Naturelle.<br />

Heiniger, U. & Rigling, D. (1994). Biological control of<br />

chestnut blight in Europe. Annual Review of<br />

Phy<strong>to</strong>pathology, 32, 581 599.<br />

Heintz, C. E. & Niederpruem, D. J. (1970). Ultrastructure<br />

and respiration of oidia and basidiospores of<br />

Coprinus lagopus (sensu Buller). Canadian Journal of<br />

Microbiology, 16, 481 484.<br />

Heintz, C. E. & Pramer, D. (1972). Ultrastructure of<br />

nema<strong>to</strong>de-trapping fungi. Journal of Bacteriology, 110,<br />

1163 1170.<br />

Heinzkill, M. & Messner, K. (1997). The ligninolytic<br />

system of fungi. In Fungal Biotechnology, ed. T. Anke.<br />

London: Chapman & Hall, pp. 213 27.<br />

Heller, A. & Gierth, K. (2001). Cy<strong>to</strong>logical observations<br />

of the infection process by Phomopsis helianthi (Munt.-<br />

Cvet) in leaves of sunflower. Journal of Phy<strong>to</strong>pathology,<br />

149, 347 357.<br />

Hemmati, F., Pell, J. K., McCartney, H. A. & Deadman,<br />

M. L. (2001). Airborne concentration of conidia of<br />

Erynia neoaphidis above cereal fields. Mycological<br />

Research, 105, 485 489.<br />

Hemmati, F., Pell, J. K., McCartney, H. A. & Deadman,<br />

M. L. (2002). Aerodynamic diameter of conidia of<br />

Erynia neoaphidis and other<br />

en<strong>to</strong>mophthoralean fungi. Mycological Research,<br />

106, 233 238.<br />

Hemmes, D. E. (1983). Cy<strong>to</strong>logy of Phy<strong>to</strong>phthora. In<br />

Phy<strong>to</strong>phthora. Its Biology, Taxonomy, Ecology, and<br />

Pathology, ed. D. C. Erwin, S. Bartnicki-Garcia & P. H.<br />

Tsao. St. Paul: APS Press, pp. 9 40.<br />

Hendrix, F. E. & Campbell, W. A. (1973). Pythiums as<br />

plant pathogens. Annual Review of Phy<strong>to</strong>pathology, 11,<br />

77 98.<br />

Henis, Y., Ghaffar, A. & Baker, R. (1979). Fac<strong>to</strong>rs<br />

affecting suppressiveness <strong>to</strong> Rhizoc<strong>to</strong>nia solani in soil.<br />

Phy<strong>to</strong>pathology, 69, 1164 1169.<br />

Hennebert, G. L. & Weresub, L. K. (1977). Terms for<br />

states of fungi, their names and types. Mycotaxon, 6,<br />

207 211.<br />

Henricot, B. & Prior, C. (2004). Phy<strong>to</strong>phthora ramorum,<br />

the cause of sudden oak death or ramorum leaf<br />

blight and dieback. Mycologist, 18, 151 156.<br />

Hepper, C. M. (1984). Isolation and culture of VA<br />

mycorrhizal (VAM) fungi. In VA Mycorrhiza, ed. C. L.<br />

Powell & D. J. Bagyaraj. Boca Ra<strong>to</strong>n: CRC Press,<br />

pp. 95 112.<br />

Hermansen, J. E., Torp, U. & Prahm, L. P. (1978). Studies<br />

of transport of cereal mildew and rust fungi across<br />

the North Sea. Grana, 17, 41 46.<br />

Herrera-Estrella, L. & Ruiz-Herrera, J. (1983). Light<br />

responses in Phycomyces blakesleeanus. Evidence for<br />

the roles of chitin biosynthesis and breakdown.<br />

Experimental Mycology, 7, 362 369.<br />

Herrera-Estrella, L., Chavez, B. & Ruiz-Herrera, J. (1982).<br />

Presence of chi<strong>to</strong>somes in the cy<strong>to</strong>plasm of<br />

Phycomyces blakesleeanus and the synthesis of chitin<br />

microfibrils. Experimental Mycology, 6, 385 388.<br />

Hervey, A. G., Bistis, G. & Leong, I. (1978). Culture<br />

studies of single ascospore isolates of Morchella<br />

esculenta. Mycologia, 70, 1269 1274.<br />

Hesseltine, C. W. (1957). The genus Syzygites<br />

(Mucoraceae). Lloydia, 20, 228 237.<br />

Hesseltine, C. W. (1991). Zygomycetes in food fermentations.<br />

Mycologist, 5, 162 169.<br />

Hesseltine, C. W. & Anderson, P. (1956). The genus<br />

Thamnidium and a study of the formation of<br />

its zygospores. American Journal of Botany, 43,<br />

696 703.<br />

Hesseltine, C. W. & Rogers, R. (1987). Dark-period<br />

induction of zygospores in Mucor. Mycologia, 79,<br />

289 297.


746 REFERENCES<br />

Hesseltine, C. W., Whitehill, A. R., Pidacks, C., et al.<br />

(1953). Coprogen, a new growth fac<strong>to</strong>r present in<br />

dung required by Pilobolus. Mycologia, 45, 7 19.<br />

Hesseltine, C. W., Benjamin, R. K. & Mehrotra, B. S.<br />

(1959). The genus Zygorhynchus. Mycologia, 51,<br />

173 194.<br />

Hibbett, D. S. & Binder, M. (2002). Evolution of complex<br />

fruiting-body morphologies in homobasidiomycetes.<br />

Proceedings of the Royal Society B, 269, 1963 1969.<br />

Hibbett, D. S. & Thorn, R. G. (1994). Nema<strong>to</strong>de-trapping<br />

in Pleurotus tuberregium. Mycologia, 86, 696 699.<br />

Hibbett, D. S. & Thorn, R. G. (2001). Basidiomycota.<br />

Homobasidiomycetes. In The Mycota VIIB: Systematics<br />

and Evolution, ed. D. J. McLaughlin, E. G. McLaughlin &<br />

P. A. Lemke. Berlin: Springer-Verlag, pp. 121 68.<br />

Hibbett, D. S., Tsuneda, A. & Murakami, S. (1994). The<br />

secotioid form of Lentinus tigrinus: genetics and<br />

development of a fungal morphological innovation.<br />

American Journal of Botany, 81, 466 478.<br />

Hibbett, D. S., Grimaldi, D. & Donoghue, M. J. (1997a).<br />

Fossil mushrooms from Miocene and Cretaceous<br />

ambers and the evolution of homobasidiomycetes.<br />

American Journal of Botany, 84, 981 991.<br />

Hibbett, D. S., Pine, E. M., Langer, E., Langer, G. &<br />

Donoghue, M. J. (1997b). Evolution of gilled mushrooms<br />

and puffballs inferred from ribosomal DNA<br />

sequences. Proceedings of the National Academy of<br />

Sciences of the USA, 94, 12 002 12006.<br />

Hicks, J. B., Strathern, J. N. & Herskowitz, I. (1977). The<br />

cassette model of mating-type interconversion. In<br />

DNA Insertion Elements, Plasmids and Episomes, ed. A. I.<br />

Bukhari, J. A. Shapiro & S. L. Adhya. Cold Spring<br />

Harbor: Cold Spring Harbor Labora<strong>to</strong>ry Press,<br />

pp. 457 462.<br />

Higham, M. T. & Cole, K. M. (1982). Fine structure of<br />

sporangiole development in Choanephora cucurbitarum<br />

(Mucorales). Canadian Journal of Botany, 60,<br />

2313 2334.<br />

Hill, D. J. & Smith, D. C. (1972). Lichen physiology. XII.<br />

The ‘inhibition technique’. New Phy<strong>to</strong>logist, 71,15 30.<br />

Hillman, B. I. & Suzuki, N. (2004). Viruses of the<br />

chestnut blight fungus, Cryphonectria parasitica.<br />

Advances in Virus Research, 63, 423 472.<br />

Hintikka, V. (1973). A note on the polarity of<br />

Armillariella mellea. Karstenia, 13, 32 39.<br />

Hin<strong>to</strong>n, D. M. & Bacon, C. W. (1985). The distribution<br />

and ultrastructure of the endophyte of <strong>to</strong>xic tall<br />

fescue. Canadian Journal of Botany, 63, 36 42.<br />

Hippe-Sanwald, S., Hermanns, M. & Somerville, S. C.<br />

(1992). Ultrastructural comparison of incompatible<br />

and compatible interactions in the barley powdery<br />

mildew disease. Pro<strong>to</strong>plasma, 168, 27 40.<br />

Hirai, A., Kano, R., Nakamura, Y., Wantanabe, S. &<br />

Hasegawa, A. (2003). Molecular taxonomy of derma<strong>to</strong>phytes<br />

and related fungi by chitin synthase 1<br />

(CHS1) gene sequences. An<strong>to</strong>nie van Leeuwenhoek, 83,<br />

11 20.<br />

Hirst, J. M. (1953). Changes in atmospheric spore<br />

content: diurnal periodicity and the effects of<br />

weather. Transactions of the British Mycological Society,<br />

36, 375 393.<br />

Hirst, J. M. & Stedman, O. J. (1960). The epidemiology of<br />

Phy<strong>to</strong>phthora infestans. II. The source of inoculum.<br />

Annals of Applied Biology, 48, 489 517.<br />

Hiruki, C. (1994). Multiple transmission of plant viruses<br />

by Olpidium brassicae. Canadian Journal of Plant<br />

Pathology, 16, 261 265.<br />

Hiura, U. (1978). Genetic basis of formae speciales in<br />

Erysiphe graminis DC. In The Powdery Mildews,<br />

ed. D. M. Spencer. London: Academic Press,<br />

pp. 101 128.<br />

Ho, H.-M. & Chen, Z.-C. (1998). Ultrastructural study of<br />

wall on<strong>to</strong>geny during zygosporogenesis in Rhizopus<br />

s<strong>to</strong>lonifer (Mucoraceae), an amended model. Botanical<br />

Bulletin of Academia Sinica, 39, 269 277.<br />

Ho, M. H.-M., Castañeda, R. F., Dugan, F. M. & Jong, S. C.<br />

(1999). Cladosporium and Cladophialophora in culture:<br />

descriptions and an expanded key. Mycotaxon, 72,<br />

115 157.<br />

Hobbie, E. A., Weber, N. S. & Trappe, J. M. (2001).<br />

Mycorrhizal vs. saprotrophic status of fungi:<br />

the iso<strong>to</strong>pic evidence. New Phy<strong>to</strong>logist, 150, 601 610.<br />

Hoch, H. C., Staples, R. C., Whitehead, B., Comeau, J. &<br />

Wolf, E. D. (1987). Signaling for growth orientation<br />

and cell differentiation by surface <strong>to</strong>pography in<br />

Uromyces. Science, 235, 1659 1662.<br />

Hock, B., Bahr, M., Walk, R.-A. & Nitschke, U. (1978). The<br />

control of fruiting body formation in the ascomycete<br />

Sordaria macrospora Auersw. by regulation of hyphal<br />

development. Planta, 141, 93 103.<br />

Hocking, D. (1963). b-Carotene and sexuality in the<br />

Mucoraceae. Nature, 197, 404.<br />

Hodge, K. T. (2003). Clavicipitaceous anamorphs. In<br />

Clavicipitalean <strong>Fungi</strong>, ed. J. F. White, C. W. Bacon, N. L.<br />

Hywel-Jones & J. W. Spatafora. New York: Marcel<br />

Dekker Inc., pp. 75 123.<br />

Hodge, K. T., Krasnoff, S. B. & Humber, R. A. (1996).<br />

Tolypocladium inflatum is the anamorph of Cordyceps<br />

subsessilis. Mycologia, 88, 715 719.<br />

Hofmann, A. (1979). LSD Mein Sorgenkind. Stuttgart:<br />

J. G. Cotta.<br />

Hogan, L. H., Klein, B. S. & Levitz, S. M. (1996). Virulence<br />

fac<strong>to</strong>rs of medically important fungi. Clinical<br />

Microbiology Reviews, 9, 469 488.


REFERENCES<br />

747<br />

Hohl, H. R. & Iselin, K. (1984). Strains of Phy<strong>to</strong>phthora<br />

infestans from Switzerland with A 2 mating type<br />

behaviour. Transactions of the British Mycological Society,<br />

83, 529 530.<br />

Hohmeyer, H. (1986). Ein Schlüssel zu den europäischen<br />

Arten der Gattung Peziza L. Zeitschrift für<br />

Mykologie, 52, 161 188.<br />

Hohn, T. M., Lovett, J. S. & Bracker, C. E. (1984).<br />

Characterization of the major proteins in gamma<br />

particles, cy<strong>to</strong>plasmic organelles in Blas<strong>to</strong>cladiella<br />

emersonii zoospores. Journal of Bacteriology, 158,<br />

253 263.<br />

Holdenrieder, O. & Greig, B. J. W. (1998). Biological<br />

methods of control. In Heterobasidion annosum.<br />

Biology, Ecology, Impact and Control, ed. S. Woodward, J.<br />

Stenlid, R. Karjalainan & A. Hüttermann.<br />

Wallingford, UK: CAB International, pp. 235 258.<br />

Holfeld, H. (1998). Fungal infections of the phy<strong>to</strong>plank<strong>to</strong>n:<br />

seasonality, minimal host density, and<br />

specificity in a mesotrophic lake. New Phy<strong>to</strong>logist, 138,<br />

507 517.<br />

Hölker, U., Érsek, T. & Hofer, M. (1993). Changes in ion<br />

fluxes and the energy demand during zoospore<br />

development in Phy<strong>to</strong>phthora infestans zoospores. Folia<br />

Microbiologica, 38, 193 200.<br />

Holliday, P. (1998). A Dictionary of Plant Pathology, 2nd<br />

edn. Cambridge: Cambridge University Press.<br />

Holliday, R. (1962). Mutation and replication in Ustilago<br />

maydis. Genetical Research, 3, 472 486.<br />

Holliday, R. (1964). A mechanism for gene conversion<br />

in fungi. Genetical Research, 5, 282 304.<br />

Hollomon, D. W. & Schmidt, H.-H. (1995). 2-<br />

Aminopyrimidine fungicides. In Modern Selective<br />

<strong>Fungi</strong>cides. Properties, Applications, Mechanisms of Action,<br />

2nd edn, ed. H. Lyr. New York: Gustav Fischer Verlag,<br />

pp. 355 371.<br />

Hollomon, D. W. & Wheeler, I. E. (2002). Controlling<br />

powdery mildews with chemistry. In The Powdery<br />

Mildews. A Comprehensive Treatise, ed. R. R. Bélanger,<br />

W. R. Bushnell, A. J. Dik & T. L. W. Carver. St. Paul,<br />

USA: APS Press, pp. 249 255.<br />

Holloway, S. A. & Heath, I. B. (1977a). Morphogenesis<br />

and the role of microtubules in synchronous<br />

populations of Saprolegnia zoospores. Experimental<br />

Mycology, 1, 9 29.<br />

Holloway, S. A. & Heath, I. B. (1977b). An ultrastructural<br />

analysis of changes in organelle arrangement and<br />

structure between the various spore types of<br />

Saprolegnia. Canadian Journal of Botany, 55,<br />

1328 1339.<br />

Holm, L. (1959). Some comments on the ascocarps of<br />

the Pyrenomycetes. Mycologia, 50, 777 788.<br />

Holst-Jensen, A., Kohn, L. M. & Schumacher, T. (1997).<br />

Nuclear rDNA phylogeny of the Sclerotiniaceae.<br />

Mycologia, 89, 885 899.<br />

Hol<strong>to</strong>rf, S., Ludwig-Müller, J., Apel, & Bohlmann, H.<br />

(1998). High level expression of a visco<strong>to</strong>xin in<br />

Arabidopsis thaliana gives enhanced resistance<br />

against Plasmodiophora brassicae. Plant Molecular<br />

Biology, 36, 673 680.<br />

Holtz, B. A., Michailides, T. J. & Hong, C. (1998).<br />

Development of apothecia from s<strong>to</strong>ne fruit infected<br />

and stromatized by Monilinia fructicola in California.<br />

Plant Disease, 82, 1375 1380.<br />

Holub, E. B., Brose, E., Tor, M., Clay, C., Crute, I. R. &<br />

Beynon, J. L. (1995). Phenotypic and genotypic variation<br />

in the interaction between Arabidopsis thaliana<br />

and Albugo candida. Molecular Plant Microbe<br />

Interactions, 8, 916 928.<br />

Honegger, R. (1978). The ascus apex in lichenized fungi<br />

I. The Lecanora-, Peltigera- and Teleoschistes-types.<br />

Lichenologist, 10, 47 67.<br />

Honegger, R. (1986). Ultrastructural studies in lichens.<br />

I. Haus<strong>to</strong>rial types and their frequencies in a range<br />

of lichens with trebouxioid phycobionts. New<br />

Phy<strong>to</strong>logist, 103, 785 795.<br />

Honegger, R. (1993). Developmental biology of lichens.<br />

New Phy<strong>to</strong>logist, 125, 659 677.<br />

Honegger, R. (1997). Metabolic interactions at the<br />

mycobiont-pho<strong>to</strong>biont interface in lichens. In The<br />

Mycota VA: Plant Relationships, ed. G. C. Carroll & P.<br />

Tudzynski. Berlin: Springer-Verlag, pp. 209 221.<br />

Honegger, R. (2001). The symbiotic phenotype of lichenforming<br />

ascomycetes. In The Mycota IX: Fungal<br />

Associations, ed. B. Hock. Berlin: Springer-Verlag,<br />

pp. 165 188.<br />

Hood, M. E. (2002). Dimorphic mating-type chromosomes<br />

in the fungus Microbotryum violaceum. Genetics,<br />

160, 457 461.<br />

Hooker, A. L. (1985). Corn and sorghum rusts. In The<br />

Cereal Rusts 2: Diseases, Distribution, Epidemiology, and<br />

Control, ed. A. P. Roelfs & W. R. Bushnell. Orlando:<br />

Academic Press, pp. 207 236.<br />

Hopple, J. S. & Vilgalys, R. (1999). Phylogenetic relationships<br />

in the mushroom genus Coprinus and darkspored<br />

allies based on sequence data from the<br />

nuclear gene coding for the large ribosomal<br />

subunit RNA: divergent domains, outgroups and<br />

monophyly. Molecular Phylogenetics and Evolution,<br />

13, 1 19.<br />

Horák, J. (2003). The role of ubiquitin in downregulation<br />

and intracellular sorting of membrane<br />

proteins: insights from yeast. Biochimica et Biophysica<br />

Acta, 1614, 139 155.


748 REFERENCES<br />

Horan, D. P., Chilvers, G. A. & Lapeyrie, F. F. (1988). Time<br />

sequence of the infection process in eucalypt<br />

ec<strong>to</strong>mycorrhizas. New Phy<strong>to</strong>logist, 109, 451 458.<br />

Horgen, P. A. & Griffin, D. H. (1969). Structure and<br />

germination of Blas<strong>to</strong>cladiella emersonii resistant<br />

sporangia. American Journal of Botany, 56, 22 25.<br />

Horgen, P. A., Meyer, R. J., Franklin, A. L., Anderson, J. B.<br />

& Filion, W. G. (1985). Motile spores from resistant<br />

sporangia of Blas<strong>to</strong>cladiella emersonii possess one-half<br />

the DNA of spores from ordinary colourless sporangia.<br />

Experimental Mycology, 9, 70 73.<br />

Horio, T. & Oakley, B. R. (2005). The role of microtubules<br />

in rapid hyphal tip growth of Aspergillus<br />

nidulans. Molecular Biology of the Cell, 16, 918 926.<br />

Horn, B. W. (1989a). Ultrastructural changes in<br />

trichospores of Smittium culisetae and S. culicis during<br />

in vitro sporangiospore extrusion and holdfast<br />

formation. Mycologia, 81, 742 753.<br />

Horn, B. W. (1989b). Requirement for potassium and pH<br />

shift in host-mediated sporangiospore extrusion<br />

from trichospores of Smittium culisetae and other<br />

Smittium species. Mycological Research, 93, 303 313.<br />

Horn, B. W. (1990). Physiological changes associated<br />

with sporangiospore extrusion from trichospores of<br />

Smittium culisetae. Experimental Mycology, 14, 113 123.<br />

Horner, J. & Moore, D. (1987). Cystidial morphogenetic<br />

field in the hymenium of Coprinus cinereus. Transactions<br />

of the British Mycological Society, 88, 479 488.<br />

Hornsey, I. S. (2003). A His<strong>to</strong>ry of Beer and Brewing.<br />

Cambridge: Royal Society of Chemistry.<br />

Horré, R. & de Hoog, G. S. (1999). Primary cerebral<br />

infections caused by melanized fungi: a review.<br />

Studies in Mycology, 43, 176 193.<br />

Horsch, M., Mayer, C., Sennhauser, U. & Rast, D. M.<br />

(1997). b-N-acetylhexosaminidase: a target for the<br />

design of antifungal agents. Pharmacology Therapy, 76,<br />

187 218.<br />

Hoshino, T., Kiriaki, M., Ohgiya, S., et al. (2003).<br />

Antifreeze proteins from snow mold fungi. Canadian<br />

Journal of Botany, 81, 1175 1181.<br />

Hostinová, E. (2002). Amylolytic enzymes produced by<br />

the yeast Saccharomycopsis fibuligera. Biologia, 57,<br />

Supplement 11, 247 251.<br />

Howard, D. H., Weitzman, I. & Padhye, A. A. (2003).<br />

Onygenales: Arthrodermataceae. In Pathogenic <strong>Fungi</strong><br />

in Humans and Animals 2nd edn, ed. D. H. Howard.<br />

New York: Marcel Dekker, pp. 141 194.<br />

Howard, K. L. (1971). Oospore types in the<br />

Saprolegniaceae. Mycologia, 63, 679 686.<br />

Howard, K. L. & Bigelow, H. E. (1969). Nutritional<br />

studies on two Gasteromycetes: Phallus ravenelii and<br />

Crucibulum levis. Mycologia, 61, 606 613.<br />

Howard, R. J. (1981). Ultrastructural analysis of hyphal<br />

tip cell growth in fungi: Spitzenkörper, cy<strong>to</strong>skele<strong>to</strong>n<br />

and endomembranes after freeze-substitution.<br />

Journal of Cell Science, 48, 89 103.<br />

Howard, R. J. (1994). Cell biology of pathogenesis. In Rice<br />

Blast Disease, ed. R. S. Zeiger, S. Leong & P. S. Teng.<br />

Wallingford, UK: CAB International, pp. 3 22.<br />

Howard, R. J. (1997). Breaching the outer barriers<br />

cuticle and cell wall penetration. In The Mycota VA:<br />

Plant Relationships, ed. G. C. Carroll & P. Tudzynski.<br />

Berlin: Springer-Verlag, pp. 43 60.<br />

Howard, R. J. & Aist, J. R. (1977). Effects of MBC on<br />

hyphal tip organization, growth and mi<strong>to</strong>sis of<br />

Fusarium acuminatum, and their antagonism by D 2 O.<br />

Pro<strong>to</strong>plasma, 92, 195 210.<br />

Howard, R. J. & Aist, J. R. (1980). Cy<strong>to</strong>plasmic microtubules<br />

and fungal morphogenesis: ultrastructural<br />

effects of methyl benzimidazole-2-ylcarbamate<br />

determined by freeze-substitution of hyphal tip cells.<br />

Journal of Cell Biology, 87, 55 64.<br />

Howard, R. J. & Valent, B. (1996). Breaking and entering:<br />

host penetration by the fungal rice blast pathogen<br />

Magnaporthe grisea. Annual Review of Microbiology, 50,<br />

491 512.<br />

Howard, R. J., Ferrari, M. A., Roach, D. H. & Money, N. P.<br />

(1991). Penetration of hard substrates by a fungus<br />

employing enormous turgor pressures. Proceedings of<br />

the National Academy of Sciences of the USA, 88,<br />

11 281 284.<br />

Howlett, B. J., Idnurm, A. & Pedras, M. S. C. (2001).<br />

Lep<strong>to</strong>sphaeria maculans, the causal agent of blackleg<br />

disease of Brassicas. Fungal Genetics and Biology, 33,<br />

1 14.<br />

Hsam, S. L. K. & Zeller, F. J. (2002). Breeding for powdery<br />

mildew resistance in common wheat (Triticum<br />

aestivum L.). In The Powdery Mildews. A Comprehensive<br />

Treatise, ed. R. R. Bélanger, W. R. Bushnell, A. J. Dik<br />

& T. L. W. Carver. St. Paul, USA: APS Press,<br />

pp. 219 38.<br />

Hu, F.-M., Zheng, R.-Y. &Chen, G.-Q. (1989). A redelimitation<br />

of the species of Pilobolus. Mycosystema, 2,<br />

111 133.<br />

Hu, G. G., Linning, R. & Bakkeren, G. (2002). Sporidial<br />

mating and infection process of the smut fungus,<br />

Ustilago hordei, in susceptible barley. Canadian Journal<br />

of Botany, 80, 1103 1114.<br />

Hu, G. G., Linning, R. & Bakkeren, G. (2003).<br />

Ultrastructural comparison of a compatible and<br />

incompatible interaction triggered by the presence<br />

of an avirulence gene during early infection of the<br />

smut fungus, Ustilago hordei, in barley. Physiological<br />

and Molecular Plant Pathology, 62, 155 166.


REFERENCES<br />

749<br />

Huang, B., Li, Z. G., Fan, M. Z. & Li, Z. Z. (2002).<br />

Molecular identification of the teleomorph of<br />

Beauveria bassiana. Mycotaxon, 81, 229 236.<br />

Huang, H.-Q. & Nielsen, J. (1984). Hybridization of the<br />

seedling-infecting Ustilago spp. pathogenic on barley<br />

and oats, and a study of the genotypes conditioning<br />

the morphology of their spore walls. Canadian Journal<br />

of Botany, 62, 603 608.<br />

Hua-Van, A., Langin, T. & Daboussi, M.-J. (2001).<br />

Evolutionary his<strong>to</strong>ry of the impala transposon in<br />

Fusarium oxysporum. Molecular Biology and Evolution, 18,<br />

1959 1969.<br />

Hubalek, Z. (2000). Keratinolytic fungi associated with<br />

free-living mammals and birds. In Biology of<br />

Derma<strong>to</strong>phytes and Other Keratinolytic <strong>Fungi</strong>, ed. R. K. S.<br />

Kushwara & J. Guarro. Bilbao: Revista<br />

Iberoamericana de Micología, pp. 93 103.<br />

Hube, B. & Naglik, J. (2001). Candida albicans proteinases:<br />

resolving the mystery of a gene family.<br />

Microbiology, 147, 1997 2005.<br />

Hudson, R. E., Aukema, J. E., Rispe, C. & Roze, D. (2002).<br />

Altruism, cheating, and anticheater adaptations<br />

in cellular slime molds. American Naturalist, 160,<br />

31 43.<br />

Hudspeth, D. S. S., Nadler, S. A. & Hudspeth, M. E. S.<br />

(2000). A COX2 molecular phylogeny of the<br />

Peronosporomycetes. Mycologia, 92, 674 684.<br />

Hughes, K. W., McGhee, L. L., Methven, A. S., Johnson,<br />

J. E. & Petersen, R. H. (1999). Patterns of geographic<br />

speciation in the genus Flammulina based on<br />

sequences of the ribosomal ITS 5.8S ITS2 area.<br />

Mycologia, 91, 978 986.<br />

Hughes, S. J. (1953). Conidiophores, conidia and classification.<br />

Canadian Journal of Botany, 31, 577 659.<br />

Hughes, S. J. (1985). The term chlamydospore. In<br />

Filamen<strong>to</strong>us Micro-Organisms. Biomedical Aspects, ed. T.<br />

Arai. Tokyo: Japanese Scientific Societies Press,<br />

pp. 1 20.<br />

Huhndorf, S. M., Miller, A. N. & Fernández, F. A. (2004).<br />

Molecular systematics of the Sordariales: the order<br />

and the family Lasiosphaeriaceae redefined.<br />

Mycologia, 96, 368 387.<br />

Humber, R. A. (1989). Synopsis of a revised classification<br />

for the En<strong>to</strong>mophthorales (Zygomycotina).<br />

Mycotaxon, 34, 441 460.<br />

Humber, R. A. (1997). <strong>Fungi</strong>: Identification. In Manual of<br />

Techniques in Insect Pathology, ed. L. A. Lacey. San Diego:<br />

Academic Press, pp. 153 185.<br />

Humber, R. A. (2000). Fungal pathogens and parasites<br />

of insects. In Applied Microbial Systematics, ed. F. Priest<br />

& M. Goodfellow. London: Chapman & Hall,<br />

pp. 199 230.<br />

Humpert, A. J., Muench, E. L., Giachini, A. J., Castellano,<br />

M. A. & Spatafora, J. W. (2001). Molecular phylogenetics<br />

of Ramaria and related genera: evidence from<br />

nuclear large subunit and mi<strong>to</strong>chondrial small<br />

subunit rDNA sequences. Mycologia, 93, 465 477.<br />

Hung, C.-Y. (1977). Ultrastructural studies of ascospore<br />

liberation in Pyronema domesticum. Canadian Journal of<br />

Botany, 55, 2544 2549.<br />

Hung, C.-Y. & Wells, K. (1971). Light and electron<br />

microscopic studies of crozier development in<br />

Pyronema domesticum. Journal of General Microbiology,<br />

66, 15 27.<br />

Hunsley, D. & Burnett, J. H. (1970). The ultrastructural<br />

architecture of the walls of some hyphal fungi.<br />

Journal of General Microbiology, 62, 203 218.<br />

Hutchison, L. J. & Kawchuk, L. M. (1998). Spongospora<br />

subterranea f. sp. subterranea. Canadian Journal of Plant<br />

Pathology, 20, 118 119.<br />

Hyde, G. J. & Hardham, A. R. (1992). Confocal microscopy<br />

of microtubule arrays in cryosectioned sporangia<br />

of Phy<strong>to</strong>phthora cinnamomi. Experimental<br />

Mycology, 16, 207 218.<br />

Hyde, G. J. & Hardham, A. R. (1993). Microtubules<br />

regulate the generation of polarity in zoospores of<br />

Phy<strong>to</strong>phthora cinnamomi. European Journal of Cell<br />

Biology, 62, 75 85.<br />

Hyde, G. J. & Heath, I. B. (1997). Ca 2þ gradients in<br />

hyphae and branches of Saprolegnia ferax. Fungal<br />

Genetics and Biology, 21, 238 251.<br />

Hyde, G. J., Gubler, F. & Hardham, A. R. (1991).<br />

Ultrastructure of zoosporogenesis in Phy<strong>to</strong>phthora<br />

cinnamomi. Mycological Research, 95, 577 591.<br />

Hyde, K. D. & Jones, E. B. G. (1989). Observations on<br />

ascospore morphology in marine fungi and their<br />

attachment <strong>to</strong> surfaces. Botanica Marina, 32,<br />

205 208.<br />

Hyde, K. D., Moss, S. T. & Jones, E. B. G. (1989).<br />

Attachment studies in marine fungi. Biofouling, 1,<br />

287 298.<br />

Hywel-Jones, N. L. (2002). Multiples of eight in Cordyceps<br />

ascospores. Mycological Research, 106, 2 3.<br />

Hywel-Jones, N. & Webster, J. (1986a). Scanning<br />

electron microscope study of external development<br />

of Erynia conica on Simulium. Transactions of the British<br />

Mycological Society, 86, 393 399.<br />

Hywel-Jones, N. & Webster, J. (1986b). Mode of infection<br />

of Simulium by Erynia conica. Transactions of the British<br />

Mycological Society, 87, 381 387.<br />

Ikediugwu, F. E. O. & Webster, J. (1970a). Antagonism<br />

between Coprinus heptemerus and other coprophilous<br />

fungi. Transactions of the British Mycological Society, 54,<br />

181 204.


750 REFERENCES<br />

Ikediugwu, F. E. O. & Webster, J. (1970b). Hyphal<br />

interference in a range of coprophilous fungi.<br />

Transactions of the British Mycological Society, 54,<br />

205 210.<br />

Ikediugwu, F. E. O., Dennis, C. & Webster, J. (1970).<br />

Hyphal interference by Peniophora gigantea against<br />

Heterobasidion annosum. Transactions of the British<br />

Mycological Society, 54, 307 309.<br />

Imai, Y. & Yamamo<strong>to</strong>, M. (1994). The fission yeast<br />

mating pheromone P-fac<strong>to</strong>r: its molecular structure,<br />

gene structure, and ability <strong>to</strong> induce gene expression<br />

and G1 arrest in the mating partner. Genes and<br />

Development, 8, 328 338.<br />

Inácio, J., Rodrigues, M. G., Sobral, P. & Fonseca, Á.<br />

(2004). Characterisation and classification of phylloplane<br />

yeasts from Portugal related <strong>to</strong> the genus<br />

Taphrina and description of five novel Lalaria species.<br />

FEMS Yeast Research, 4, 541 555.<br />

Inamadar, A. C. & Palit, A. (2003). The genus Malassezia<br />

and human disease. Indian Journal of Derma<strong>to</strong>logy,<br />

Venereology and Leprology, 69, 265 270.<br />

Ing, B. (1998). Exobasidium in the British Isles. Mycologist,<br />

12, 80 82.<br />

Ing, B. (1999). The Myxomycetes of Britain and Ireland.<br />

Slough: Richmond Publishing Company.<br />

Inglis, G. D., Yanke, L. J., Kawchuk, L. M. & McAllister,<br />

T. A. (1999). The influence of bacterial inoculants on<br />

the microbial ecology of aerobic spoilage of barley<br />

silage. Canadian Journal of Microbiology, 45, 77 87.<br />

Inglis, R. J., Langan, T. A., Matthews, H. R., Hardie, D. G.<br />

& Bradbury, E. M. (1976). Advance of mi<strong>to</strong>sis by<br />

his<strong>to</strong>ne phosphokinase. Experimental Cell Research, 97,<br />

418 425.<br />

Ingold, C. T. (1939). Spore Discharge in Land Plants.<br />

Oxford: Clarendon Press.<br />

Ingold, C. T. (1946). Spore discharge in Daldinia concentrica.<br />

Transactions of the British Mycological Society, 29,<br />

43 51.<br />

Ingold, C. T. (1953). Dispersal in <strong>Fungi</strong>. Oxford: Clarendon<br />

Press.<br />

Ingold, C. T. (1954a). Ascospore form. Transactions of the<br />

British Mycological Society, 37, 19 21.<br />

Ingold, C. T. (1954b). The ascogenous hyphae in Daldinia<br />

concentrica. Transactions of the British Mycological Society,<br />

37, 108 110.<br />

Ingold, C. T. (1956). A gas phase in viable fungal spores.<br />

Nature, 177, 1242 1243.<br />

Ingold, C. T. (1957). Spore liberation in higher fungi.<br />

Endeavour, 16, 78 83.<br />

Ingold, C. T. (1959). Jelly as a water reserve in fungi.<br />

Transactions of the British Mycological Society, 42,<br />

475 478.<br />

Ingold, C. T. (1966). Aspects of spore liberation: violent<br />

discharge. In The Fungus Spore, ed. M. F. Madelin.<br />

London: Butterworths, pp. 113 32.<br />

Ingold, C. T. (1971). Fungal Spores. Their Liberation and<br />

Dispersal. Oxford: Clarendon Press.<br />

Ingold, C. T. (1972). Sphaerobolus: the s<strong>to</strong>ry of a fungus.<br />

Transactions of the British Mycological Society, 58,<br />

179 195.<br />

Ingold, C. T. (1975). An Illustrated Guide <strong>to</strong> Aquatic and<br />

Water-Borne Hyphomycetes (<strong>Fungi</strong> Imperfecti) with Notes<br />

on their Biology. Ambleside, UK: Freshwater Biological<br />

Association.<br />

Ingold, C. T. (1980). Mycelia, oidia and sporophore<br />

initials in Flammulina velutipes. Transactions of the<br />

British Mycological Society, 75, 107 116.<br />

Ingold, C. T. (1981). The first-formed phialoconidium of<br />

Thielaviopsis basicola. Transactions of the British<br />

Mycological Society, 76, 517 519.<br />

Ingold, C. T. (1982a). Basidiospore germination and<br />

conidium development in Auricularia.<br />

Transactions of the British Mycological Society, 78,<br />

161 166.<br />

Ingold, C. T. (1982b). Basidiospore germination and<br />

conidium formation in Exidia glandulosa and Tremella<br />

mesenterica. Transactions of the British Mycological<br />

Society, 79, 370 373.<br />

Ingold, C. T. (1983a). Structure and development in an<br />

isolate of Itersonilia perplexans. Transactions of the<br />

British Mycological Society, 80, 365 368.<br />

Ingold, C. T. (1983b). Basidiospore germination and<br />

conidium development in Dacrymycetales.<br />

Transactions of the British Mycological Society, 81,<br />

563 571.<br />

Ingold, C. T. (1983c). The basidium in Ustilago.<br />

Transactions of the British Mycological Society, 81,<br />

573 584.<br />

Ingold, C. T. (1984a). Further observations on Itersonilia.<br />

Transactions of the British Mycological Society, 83,<br />

166 174.<br />

Ingold, C. T. (1984b). Patterns of ballis<strong>to</strong>spore germination<br />

in Tilletiopsis, Auricularia and Tulasnella.<br />

Transactions of the British Mycological Society, 83,<br />

583 591.<br />

Ingold, C. T. (1985). Observations on spores and their<br />

germination in certain Heterobasidiomycetes.<br />

Transactions of the British Mycological Society, 85,<br />

417 423.<br />

Ingold, C. T. & Cox, V. J. (1955). Periodicity of spore<br />

discharge in Daldinia concentrica. Annals of Botany N. S.,<br />

21, 201 209.<br />

Ingold, C. T. & Hadland, S. A. (1959). The ballistics of<br />

Sordaria. New Phy<strong>to</strong>logist, 58, 46 57.


REFERENCES<br />

751<br />

Ingold, C. T. & Zoberi, M. H. (1963). The asexual<br />

apparatus of Mucorales in relation <strong>to</strong> spore liberation.<br />

Transactions of the British Mycological Society, 46,<br />

115 134.<br />

Ingold, C. T., Davey, R. A. & Wakley, G. (1981). The<br />

teliospore pedicel of Phragmidium mucronatum.<br />

Transactions of the British Mycological Society, 77,<br />

439 442.<br />

Ingram, D. S. & Williams, P. H., eds. (1991). Phy<strong>to</strong>phthora<br />

infestans, the Cause of Late Blight of Pota<strong>to</strong>. London:<br />

Academic Press.<br />

Innes, J. L. (1988). The use of lichens in dating. In CRC<br />

Handbook of Lichenology III, ed. M. Galun. Boca Ra<strong>to</strong>n:<br />

CRC Press, pp. 75 91.<br />

Innocenti, G., Roberti, R., Montanari, M. & Zakrisson, E.<br />

(2003). Efficacy of microorganisms antagonistic <strong>to</strong><br />

Rhizoc<strong>to</strong>nia cerealis and their cell wall degrading<br />

enzymatic activities. Mycological Research, 107,<br />

421 427.<br />

Iqbal, S. H. & Webster, J. (1973a). The trapping of<br />

aquatic hyphomycete spores by air bubbles.<br />

Transactions of the British Mycological Society, 60, 37 48.<br />

Iqbal, S. H. & Webster, J. (1973b). Aquatic hyphomycete<br />

spora of the River Exe and its tributaries. Transactions<br />

of the British Mycological Society, 61, 331 346.<br />

Isaka, M., Kittakoop, P. & Thebtaranonth, Y. (2003).<br />

Secondary metabolites of Clavicipitalean fungi. In<br />

Clavicipitalean <strong>Fungi</strong>, ed. J. F. White, C. W. Bacon, N. L.<br />

Hywel-Jones & J. W. Spatafora. New York: Marcel<br />

Dekker Inc., pp. 355 397.<br />

Ishibashi, Y., Sakagami, Y., Isogai, A. & Suzuki, A. (1984).<br />

Structures of tremerogens A-9291 I and A-<br />

9291 VIII: peptidyl sex hormones of Tremella brasiliensis.<br />

Biochemistry, 23, 1399 1404.<br />

Issakainen, J., Jalava, J., Hyvönen, J., et al. (2003).<br />

Relationships of Scopulariopsis based on LSU rDNA<br />

sequences. Medical Mycology, 41, 31 42.<br />

Iten, W. & Matile, P. (1970). Role of chitinase and other<br />

lysosomal enzymes of Coprinus lagopus in the au<strong>to</strong>lysis<br />

of fruiting bodies. Journal of General Microbiology,<br />

61, 301 309.<br />

I<strong>to</strong>, H., Hanyaku, H., Harada, T., Mochizuki, T. &<br />

Tanaka, S. (1998). Ultrastructure of the<br />

ascospore formation of Arthroderma simii. Mycoses,<br />

41, 133 137.<br />

Jackson, G. V. H. & Gay, J. L. (1976). Perennation of<br />

Sphaerotheca mors-uvae as cleis<strong>to</strong>thecia with particular<br />

reference <strong>to</strong> microbial activity. Transactions of the<br />

British Mycological Society, 66, 463 471.<br />

Jaffee, B. A. (1992). Population biology and biological<br />

control of nema<strong>to</strong>des. Canadian Journal of<br />

Microbiology, 38, 359 364.<br />

Jahn, M., Munger, H. M. & McCreight, J. D. (2002).<br />

Breeding cucurbit crops for powdery mildew resistance.<br />

In The Powdery Mildews. A Comprehensive Treatise,<br />

ed. R. R. Bélanger, W. R. Bushnell, A. J. Dik &<br />

T. L. W. Carver. St. Paul, USA: APS Press, pp. 239 248.<br />

Jailloux, F., Thind, T. & Clerjeau, M. (1998). Release,<br />

germination, and pathogenicity of ascospores of<br />

Uncinula neca<strong>to</strong>r under controlled conditions.<br />

Canadian Journal of Botany, 76, 777 781.<br />

Jailloux, F., Willocquet, L., Chapuis, L. & Froidefond, G.<br />

(1999). Effect of weather fac<strong>to</strong>rs on the release of<br />

ascospores of Uncinula neca<strong>to</strong>r, the cause of grape<br />

powdery mildew, in the Bordeaux region. Canadian<br />

Journal of Botany, 77, 1044 1051.<br />

Jain, N., Li, L., McFadden, D. C., et al. (2006). Phenotypic<br />

switching in a Cryp<strong>to</strong>coccus neoformans variety gattii<br />

strain is associated with changes in virulence and<br />

promotes dissemination <strong>to</strong> the central nervous<br />

system. Infection and Immunity, 74, 896 903.<br />

Jakobsen, I. & Rosendahl, L. (1990). Carbon flow in<strong>to</strong><br />

soil and external hyphae from roots of mycorrhizal<br />

cucumber plants. New Phy<strong>to</strong>logist, 115, 77 83.<br />

Jakobsen, M., Can<strong>to</strong>r, M. D. & Jespersen, L. (2002).<br />

Production of bread, cheese and meat. In The Mycota<br />

X: Industrial Applications, ed. H. D. Osiewacz. Berlin:<br />

Springer-Verlag, pp. 3 22.<br />

James, T. Y. & Vilgalys, R. (2001). Abundance and<br />

diversity of Schizophyllum commune spore clouds in<br />

the Caribbean detected by selective sampling.<br />

Molecular Ecology, 10, 471 479.<br />

James, T. Y., Porter, D., Leander, C. A., Vilgalys, R. &<br />

Longcore, J. E. (2000). Molecular phylogenetics of the<br />

Chytridiomycota supports the utility of ultrastructural<br />

data in chytrid systematics. Canadian Journal of<br />

Botany, 78, 336 350.<br />

Janex-Favre, M. C. & Parguey-Leduc, A. (2002).<br />

Particularités des ascospores et de l’hyménium des<br />

truffes (Ascomycètes) I. Développement et<br />

structure des ascocarpes. Cryp<strong>to</strong>gamie Mycologie, 23,<br />

103 128.<br />

Janex-Favre, M.-C., Parguey-Leduc, A. & Bruxelles, G.<br />

(1998). L’hyménium de Morchella deliciosa Fr.<br />

(Ascomycètes, Discomycètes). Cryp<strong>to</strong>gamie<br />

Bryologie Lichénologie, 19, 293 304.<br />

Janisiewicz, W. J. & Bors, B. (1995). Development of a<br />

microbial community of bacterial and yeast antagonists<br />

<strong>to</strong> control wound-invading postharvest pathogens<br />

of fruits. Applied and Environmental Microbiology,<br />

61, 3261 3267.<br />

Janisiewicz, W. J. & Korsten, L. (2002). Biological control<br />

of postharvest diseases of fruits. Annual Review of<br />

Phy<strong>to</strong>pathology, 40, 411 441.


752 REFERENCES<br />

Jansson, H.-B. & Nordbring-Hertz, B. (1979). Attraction<br />

of nema<strong>to</strong>des <strong>to</strong> living mycelium of<br />

nema<strong>to</strong>phagous fungi. Journal of General Microbiology,<br />

112, 89 93.<br />

Jansson, H.-B. & Nordbring-Hertz, B. (1980). Interactions<br />

between nema<strong>to</strong>phagous fungi and plant-parasitic<br />

nema<strong>to</strong>des: attraction, induction of trap formation<br />

and capture. Nema<strong>to</strong>logica, 26, 383 389.<br />

Jansson, H.-B., Jeyaprakash, A. & Zuckerman, B. M.<br />

(1985). Differential adhesion and infection of nema<strong>to</strong>des<br />

by the endoparasitic fungus Meria coniospora<br />

(Deuteromycetes). Applied and Environmental<br />

Microbiology, 49, 552 555.<br />

Jarl, K. (1969). Symba yeast process. Food Technology, 23,<br />

23 26.<br />

Jarvis, W. R., Gubler, W. D. & Grove, G. G. (2002).<br />

Epidemiology of powdery mildews in agricultural<br />

pathosystems. In The Powdery Mildews. A Comprehensive<br />

Treatise, ed. R. R. Bélanger, W. R. Bushnell,<br />

A. J. Dik & T. L. W. Carver. St. Paul, USA: APS Press,<br />

pp. 169 199.<br />

Javadekar, V. S., Sivaraman, H. & Gokhale, D. V. (1995).<br />

Industrial yeast strain improvement: construction of<br />

a highly flocculent yeast with a killer character by<br />

pro<strong>to</strong>plast fusion. Journal of Industrial Microbiology, 15,<br />

94 102.<br />

Jazwinski, S. M. (2002). Growing old: metabolic control<br />

and yeast aging. Annual Review of Microbiology, 56,<br />

769 792.<br />

Jee, H. J., Ho, H. H. & Cho, W. D. (2000). Pythioge<strong>to</strong>n zeae<br />

sp. nov. causing root and basal stalk rot of corn in<br />

Korea. Mycologia, 92, 522 527.<br />

Jeffries, P. (1999). Scleroderma. InEc<strong>to</strong>mycorrhizal <strong>Fungi</strong>:<br />

Key Genera in Profile, ed. J. W. G. Cairney & S. M.<br />

Chambers. Berlin: Springer-Verlag, pp. 187 200.<br />

Jeffries, P. & Young, T. W. K. (1975). Ultrastructure of<br />

the sporangiospores of Pip<strong>to</strong>cephalis unispora. Archiv<br />

für Mikrobiologie, 105, 329 333.<br />

Jeffries, P. & Young, T. W. K. (1976). Ultrastructure of<br />

infection of Cokeromyces recurvatus by Pip<strong>to</strong>cephalis<br />

unispora (Mucorales). Archives of Microbiology, 109,<br />

277 288.<br />

Jeffries, P. & Young, T. W. K. (1994). Interfungal Parasitic<br />

Relationships. Wallingford, UK: CAB International.<br />

Jeffries, T. W. & Kurtzman, C. P. (1994). Strain selection,<br />

taxonomy, and genetics of xylose-fermenting yeasts.<br />

Enzyme and Microbial Technology, 16, 922 932.<br />

Jeger, M. J., Gilijamse, E., Bock, C. H. & Frinking, H. D.<br />

(1998). The epidemiology, variability and control of<br />

the downy mildews of pearl millet and sorghum,<br />

with particular reference <strong>to</strong> Africa. Plant Pathology,<br />

47, 544 569.<br />

Jeng, R. S. & Hubbes, M. (1980). Ultrastructure of<br />

Cera<strong>to</strong>cystis ulmi. II. Ascogenous system and ascosporogenesis.<br />

European Journal of Forest Pathology, 10,<br />

104 116.<br />

Jenkinson, P. & Parry, D. W. (1994). Splash dispersal of<br />

conidia of Fusarium culmorum and Fusarium avenaceum.<br />

Mycological Research, 98, 506 510.<br />

Jenkyn, J. F. & Bainbridge, A. (1978). Biology and<br />

pathology of cereal powdery mildews. In The Powdery<br />

Mildews, ed. D. M. Spencer. London: Academic Press,<br />

pp. 283 321.<br />

Jennersten, O. (1988). Insect dispersal of fungal disease:<br />

effects of Ustilago infection on pollina<strong>to</strong>r attraction<br />

in Viscaria vulgaris. Oikos, 51, 163 170.<br />

Jennings, D. H. (1987). Translocation of solutes in fungi.<br />

Biological Reviews, 62, 215 243.<br />

Jennings, D. H. (1991). The physiology and biochemistry<br />

of the vegetative mycelium. In Serpula lacrymans.<br />

Fundamental Biology and Control Strategies, ed. D. H.<br />

Jennings & A. F. Bravery. Chichester: John Wiley &<br />

Sons, pp. 55 79.<br />

Jennings, D. H. (1995). The Physiology of Fungal Nutrition.<br />

Cambridge: Cambridge University Press.<br />

Jennings, D. H. & Bravery, A. F., eds. (1991). Serpula<br />

lacrymans. Fundamental Biology and Control Strategies.<br />

Chichester: John Wiley & Sons.<br />

Jennings, D. H. & Watkinson, S. C. (1982). Structure and<br />

development of mycelial strands in Serpula lacrimans.<br />

Transactions of the British Mycological Society, 78,<br />

465 474.<br />

Jensen, A. B., Gargas, A., Eilenberg, J. & Rosendahl, S.<br />

(1998). Relationships of the insect-pathogenic order<br />

En<strong>to</strong>mophthorales (Zygomycota, <strong>Fungi</strong>) based on<br />

phylogenetic analyses of nuclear small subunit<br />

ribosomal DNA sequences (SSU rDNA). Fungal Genetics<br />

and Biology, 24, 325 334.<br />

Jensen, J. D. (1983). The development of Diaporthe<br />

phaseolorum variety sojae in culture. Mycologia, 75,<br />

1074 1091.<br />

Jensen, R. E., Hobbs, A. E. A., Cerveny, K. L. & Sesaki, H.<br />

(2000). Yeast mi<strong>to</strong>chondrial dynamics: fusion, division,<br />

segregation, and shape. Microscopy Research and<br />

Technique, 51, 573 583.<br />

Jenson, I. (1998). Bread and baker’s yeast. In Microbiology<br />

of Fermented Foods, 2nd edn, ed. B. J. B. Wood. London:<br />

Blackie Academic, pp. 172 95.<br />

Jesenská, Z., Piecková, E. & Bernát, D. (1993). Heat<br />

resistance of fungi from soil. International Journal of<br />

Food Microbiology, 19, 187 192.<br />

Ji, T. & Dayal, R. (1971). Studies in the life cycle<br />

of Allomyces javanicus Kniep. Hydrobiologia, 37,<br />

245 251.


REFERENCES<br />

753<br />

Jiang, J., Stephenson, L. W., Erwin, D. C. & Leary, J. V.<br />

(1989). Nuclear changes in Phy<strong>to</strong>phthora during<br />

oospore maturation and germination. Mycological<br />

Research, 92, 463 469.<br />

Jiang, Y. & Yao, Y.-J. (2002). Names related <strong>to</strong> Cordyceps<br />

sinensis anamorph. Mycotaxon, 84, 245 254.<br />

Joffe, A. Z. (1986). Fusarium Species: Their Biology and<br />

Toxicology. New York: John Wiley & Sons.<br />

Johannesson, H., Læssøe, T. & Stenlid, J. (2000).<br />

Molecular and morphological investigation of<br />

Daldinia in northern Europe. Mycological Research, 104,<br />

275 280.<br />

Johannesson, H., Gustafsson, M. & Stenlid, J. (2001).<br />

Local population structure of the wood decay<br />

ascomycete Daldinia loculata. Mycologia, 93, 440 446.<br />

Johnson, B. F., Sowden, L. C., Walker, T., Yoo, B. Y. &<br />

Calleja, G. B. (1989). Use of electron microscopy <strong>to</strong><br />

characterize the surfaces of flocculent and nonflocculent<br />

yeast cells. Canadian Journal of Microbiology, 35,<br />

1081 1086.<br />

Johnson, E. A. & An, G.-H. (1991). Astaxanthin from<br />

microbial sources. Critical Reviews in Biotechnology, 11,<br />

297 326.<br />

Johnson, E. A. & Schroeder, W. A. (1995a). Microbial<br />

carotenoids. Advances in Biochemical Engineering/<br />

Biotechnology, 53, 119 178.<br />

Johnson, E. A. & Schroeder, W. (1995b). Astaxanthin<br />

from the yeast Phaffia rhodozyma. Studies in Mycology,<br />

38, 81 90.<br />

Johnson, E. M. & Valleau, W. D. (1949). Synonymy in<br />

some common species of Cercospora. Phy<strong>to</strong>pathology,<br />

39, 763 770.<br />

Johnson, T. (1953). Variation in the rusts of cereals.<br />

Biological Reviews, 28, 105 157.<br />

Johnson, T. W. (1969). The aquatic fungi of Iceland:<br />

Olpidium (Braun) Rabenhorst. Archiv für Mikrobiologie,<br />

69, 1 11.<br />

Johnson, T. W. (1973). Aquatic fungi from Iceland: some<br />

polycentric species. Mycologia, 65, 1337 1355.<br />

Johnson, T. W. (1977). Resting spore germination in<br />

three chytrids. Mycologia, 69, 34 45.<br />

Johns<strong>to</strong>n, P. R. (1997). Tropical Rhytismatales. In<br />

Biodiversity of Tropical Microfungi, ed. K. D. Hyde. Hong<br />

Kong: Hong Kong University Press, pp. 241 254.<br />

Joly, P. (1964). Le Genre Alternaria. Paris: <strong>Edition</strong>s Paul<br />

Chevalier.<br />

Jones, A. M. & Jones, E. B. G. (1993). Observations on the<br />

marine gasteromycete Nia vibrissa. Mycological<br />

Research, 97, 1 6.<br />

Jones, B. E. & Gooday, G. W. (1977). Lectin binding <strong>to</strong><br />

sexual cells in fungi. Biochemical Society Transactions, 5,<br />

717 719.<br />

Jones, D., McHardy, W. J. & Wilson, M. J. (1976).<br />

Ultrastructure and chemical composition of spines<br />

in Mucorales. Transactions of the British Mycological<br />

Society, 66, 153 157.<br />

Jones, E. B. G. (1994). Fungal adhesion. Mycological<br />

Research, 98, 961 981.<br />

Jones, P. (1999). Control of loose smut (Ustilago nuda<br />

and U. tritici) infections in barley and wheat by foliar<br />

applications of systemic fungicides. European Journal<br />

of Plant Pathology, 105, 729 732.<br />

Jones, S. G. (1925). Life-his<strong>to</strong>ry and cy<strong>to</strong>logy of Rhytisma<br />

acerinum (Pers.) Fries. Annals of Botany, 39, 41 73.<br />

Jones, S. W., Donaldson, S. P. & Deacon, J. W. (1991).<br />

Behaviour of zoospores and zoospore cysts in<br />

relation <strong>to</strong> root infection by Pythium aphanidermatum.<br />

New Phy<strong>to</strong>logist, 117, 289 301.<br />

Joosten, M. H. A. J., Hendrickx, L. J. M. & de Wit, P. G. J. M.<br />

(1990). Carbohydrate composition of apoplastic<br />

fluids isolated from <strong>to</strong>ma<strong>to</strong> leaves inoculated with<br />

virulent or avirulent races of Cladosporium fulvum<br />

(syn. Fulvia fulva). Netherlands Journal of Plant Pathology,<br />

96, 103 112.<br />

Joosten, M. H. A. J., Honée, G., van Kan, J. A. L. & de Wit,<br />

P. J. G. M. (1997). The gene-for-gene concept in plantpathogen<br />

interactions: <strong>to</strong>ma<strong>to</strong> Cladosporium fulvum.<br />

In The Mycota VB: Plant Relationships, ed. G. C. Carroll &<br />

P. Tudzynski. Berlin: Springer-Verlag, pp. 3 16.<br />

Jordan, M. (1995). Evidence of severe allergic reactions<br />

<strong>to</strong> Laetiporus sulphureus. Mycologist, 9, 157 158.<br />

Jørgensen, J. H. (1992). Discovery, characterization and<br />

exploitation of Mlo powdery mildew resistance genes<br />

in barley. Euphytica, 63, 141 152.<br />

Jørgensen, J. H. (1994). Genetics of powdery mildew<br />

resistance in barley. Critical Reviews in Plant Science, 13,<br />

97 119.<br />

Ju, Y.-M. & Rogers, J. D. (1996). A Revision of the Genus<br />

Hypoxylon. St. Paul, USA: APS Press.<br />

Ju, Y.-M., Rogers, J. D. & San Martin, F. (1997). A revision<br />

of the genus Daldinia. Mycotaxon, 61, 243 293.<br />

Jurgens, J. A., Blanchette, R. A., Zambino, P. J. & David,<br />

A. (2003). His<strong>to</strong>logy of white pine blister rust in<br />

needles of resistant and susceptible Eastern white<br />

pine. Plant Disease, 87, 1026 1030.<br />

Kahmann, R., Steinberg, G., Basse, C., Feldbrügge, M. &<br />

Kämper, J. (2000). Ustilago maydis, the causative agent<br />

of corn smut disease. In Fungal Pathology, ed. J. W.<br />

Kronstad. Dordrecht: Kluwer Academic Publishers,<br />

pp. 347 371.<br />

Kakani, K., Robbins, M. & Rochon, D. (2003). Evidence<br />

that binding of cucumber necrosis virus <strong>to</strong> vec<strong>to</strong>r<br />

zoospores involves recognition of oligosaccharides.<br />

Journal of Virology, 77, 3922 3928.


754 REFERENCES<br />

Kaltz, O. & Shykoff, J. A. (2001). Male and female Silene<br />

latifolia plants differ in per-contact risk of infection<br />

by a sexually transmitted disease. Journal of Ecology,<br />

89, 99 109.<br />

Kam, A. P. & Xu, J. (2002). Diversity of commensal yeasts<br />

within and among healthy hosts. Diagnostic<br />

Microbiology and Infectious Disease, 43, 19 28.<br />

Kamada, T. (1994). Stipe elongation in fruit bodies. In<br />

The Mycota I: Growth, Differentiation and Sexuality, ed.<br />

J. G. H. Wessels & F. Meinhardt. Berlin: Springer-<br />

Verlag, pp. 367 379.<br />

Kamada, T., Hirai, K. & Fujii, M. (1993). The role of the<br />

cy<strong>to</strong>skele<strong>to</strong>n in the pairing and positioning of the<br />

two nuclei in the apical cells of the dikaryon of the<br />

basidiomycete Coprinus cinereus. Experimental<br />

Mycology, 17, 338 344.<br />

Kaminskyj, S. G. W. & Hamer, J. E. (1998). hyp loci<br />

control cell pattern formation in the vegetative<br />

mycelium of Aspergillus nidulans. Genetics, 148,<br />

669 680.<br />

Kaminskyj, S. G. W. & Heath, I. B. (1996). Studies on<br />

Saprolegnia ferax suggest the general importance of<br />

the cy<strong>to</strong>plasm in determining hyphal morphology.<br />

Mycologia, 88, 20 37.<br />

Kamiya, Y., Sakurai, A., Tamura, S., et al. (1979).<br />

Structure of rhodo<strong>to</strong>rucine A, a peptidyl fac<strong>to</strong>r,<br />

inducing mating tube formation in Rhodosporidium<br />

<strong>to</strong>ruloides. Agricultural and Biological Chemistry, 43,<br />

363 369.<br />

Kämper, J., Bölker, M. & Kahmann, R. (1994). Matingtype<br />

genes in Heterobasidiomycetes. In The Mycota I:<br />

Growth, Differentiation and Sexuality, ed. J. G. H. Wessels<br />

& F. Meinhardt. Berlin: Springer-Verlag, pp. 323 332.<br />

Kapat, A., Zimand, G. & Elad, Y. (1998). Biosynthesis of<br />

pathogenicity hydrolytic enzymes by Botrytis cinerea<br />

during infection of bean leaves and in vitro.<br />

Mycological Research, 102, 1017 1024.<br />

Kaplan, J. D. & Goos, R. D. (1982). The effect of water<br />

potential on zygospore formation in Syzygites megalocarpus.<br />

Mycologia, 74, 684 686.<br />

Kapteyn, J. C., van den Ende, H. & Klis, F. M. (1999). The<br />

contribution of yeast wall proteins <strong>to</strong> the organization<br />

of the yeast cell wall. Biochimica et Biophysica Acta,<br />

1426, 373 383.<br />

Karaffa, L. & Kubicek, C. P. (2003). Aspergillus niger citric<br />

acid accumulation: do we understand this well<br />

working black box? Applied Microbiology and<br />

Biotechnology, 61, 189 196.<br />

Karjalainen, R. & Lounatmaa, K. (1986). Ultrastructure<br />

of penetration and colonization of wheat leaves by<br />

Sep<strong>to</strong>ria nodorum. Physiological and Molecular Plant<br />

Pathology, 29, 263 270.<br />

Karling, J. S. (1950). The genus Physoderma. Lloydia, 13,<br />

1 71.<br />

Karling, J. S. (1964). Synchytrium. New York and London:<br />

Academic Press.<br />

Karling, J. S. (1968). The Plasmodiophorales. New York and<br />

London: Hafner Publishing Company.<br />

Karling, J. S. (1969). Zoosporic fungi in Oceania. VII.<br />

Fusions in Rhizophlyctis. American Journal of Botany, 56,<br />

211 221.<br />

Karling, J. S. (1973). A note on Blas<strong>to</strong>cladiella<br />

(Blas<strong>to</strong>cladiaceae). Mycopathologia et Mycologia<br />

Applicata, 49, 169 172.<br />

Karling, J. S. (1977). Chytridiomycetarum Iconographia.<br />

Monticello and New York: Lubrecht & Cramer.<br />

Kars, I. & van Kan, J. A. L. (2004). Extracellular enzymes<br />

and metabolites involved in pathogenesis of Botrytis.<br />

In Botrytis: Biology, Pathology and Control, ed. Y. Elad, B.<br />

Williamson, P. Tudzynski & N. Delen. Dordrecht:<br />

Kluwer, pp. 99 118.<br />

Kataria, H. R. & Hoffmann, G. M. (1988). A critical<br />

review of plant pathogenic species of Cera<strong>to</strong>basidium<br />

Rogers. Zeitschrift für Pflanzenkrankheiten und<br />

Pflanzenschutz, 95, 81 107.<br />

Kauppi, M. (1979). The exploitation of Cladonia stellaris<br />

in Finland. Lichenologist, 11, 85 89.<br />

Kauserud, H., Högberg, N., Knudsen, H., Elborne, S. A. &<br />

Schumacher, T. (2004). Molecular phylogenetics<br />

suggest a North American link between the anthropogenic<br />

dry rot fungus Serpula lacrymans and its wild<br />

relative S. himantioides. Molecular Ecology, 13,<br />

3137 3146.<br />

Kaushik, N. K. & Hynes, H. B. N. (1968). Experimental<br />

study on the role of autumn-shed leaves in aquatic<br />

environments. Journal of Ecology, 56, 229 243.<br />

Kaushik, N. K. & Hynes, H. B. N. (1971). The fate of dead<br />

leaves that fall in<strong>to</strong> streams. Archiv für Hydrobiologie,<br />

68, 465 515.<br />

Keenan, K. A. & Weiss, R. L. (1997). Characterization of<br />

vacuolar arginine uptake and amino acid efflux in<br />

Neurospora crassa using cupric ion <strong>to</strong> permeabilize<br />

the plasma membrane. Fungal Genetics and Biology, 22,<br />

177 190.<br />

Kemen, E., Kemen, A. C., Rafiqi, M., et al. (2005).<br />

Identification of a protein from rust fungi transferred<br />

from haus<strong>to</strong>ria in<strong>to</strong> infected plant cells.<br />

Molecular Plant Microbe Interactions, 18, 1130 1139.<br />

Kemp, R. F. O. (1975a). Breeding biology of Coprinus<br />

species in the section Lanatuli. Transactions of the<br />

British Mycological Society, 65, 375 388.<br />

Kemp, R. F. O. (1975b). Oidia, plasmogamy and speciation<br />

in basidiomycetes. Biological Journal of the Linnean<br />

Society, 7, S57 S69.


REFERENCES<br />

755<br />

Kendrick, B., ed. (1971). Taxonomy of <strong>Fungi</strong> Imperfecti.<br />

Toron<strong>to</strong> and Buffalo: Toron<strong>to</strong> University Press.<br />

Kendrick, W. B. & Cole, G. T. (1968). Conidium on<strong>to</strong>geny<br />

in Hyphomycetes. The sympodulae of Beauveria<br />

and Curvularia. Canadian Journal of Botany, 46,<br />

1297 1301.<br />

Kendrick, B. & Watling, R. (1979). Mi<strong>to</strong>spores in<br />

basidiomycetes. In The Whole Fungus 2, ed. B.<br />

Kendrick. Ottawa: National Museum of Natural<br />

Sciences, National Museums of Canada, Kananaskis<br />

Foundation, pp. 473 545.<br />

Kerkenaar, A. (1995). Mechanism of action of cyclic<br />

amine fungicides: morpholines and piperidines. In<br />

Modern Selective <strong>Fungi</strong>cides. Properties, Applications,<br />

Mechanisms of Action, 2nd edn, ed. H. Lyr. New York:<br />

Gustav Fischer Verlag, pp. 185 204.<br />

Kern, H. & Naef-Roth, S. (1975). Zur Bildung von<br />

Auxinen und Cy<strong>to</strong>kininen durch Taphrina-Arten.<br />

Phy<strong>to</strong>pathologische Zeitschrift, 83, 103 108.<br />

Kern, V. D. (1999). Gravitropism of basidiomyce<strong>to</strong>us<br />

fungi. Life Sciences: Microgravity Research, 24, 697 706.<br />

Kern, V. D., Rehm, A. & Hock, B. (1998). Gravitropic<br />

bending of fruit bodies a model based on hyphal<br />

gravisensing and cooperativity. Life Sciences:<br />

Microgravity Research, 21, 1173 1178.<br />

Kerrigan, R. W. (1995). Global genetic resources for<br />

Agaricus breeding and cultivation. Canadian Journal of<br />

Botany, 73, S973 S979.<br />

Kerrigan, R. W. (1996). Characteristics of a large<br />

collection of wild edible mushroom germplasm: the<br />

Agaricus Resource Program. In Culture Collections <strong>to</strong><br />

Improve the Quality of Life, ed. R. A. Samson,<br />

J. A. Stalpers, O. van der Mei & A. H. S<strong>to</strong>uthamer.<br />

Baarn: Centraalbureau voor Schimmelcultures,<br />

pp. 302 308.<br />

Kerrigan, R. W., Imbernon, M., Callac, P., Billette, C. &<br />

Oliver, J. M. (1994). The heterothallic life cycle of<br />

Agaricus bisporus var burnettii and the inheritance of<br />

the tetrasporic trait. Experimental Mycology, 18,<br />

193 210.<br />

Kerry, B. R. (2000). Rhizosphere interactions and the<br />

exploitation of microbial agents for the biological<br />

control of plant-parasitic nema<strong>to</strong>des. Annual Review of<br />

Phy<strong>to</strong>pathology, 38, 423 441.<br />

Kerry, B. R. & Jaffee, B. A. (1997). <strong>Fungi</strong> as biological<br />

control agents for plant parasitic nema<strong>to</strong>des. In The<br />

Mycota IV: Environmental and Microbial Relationships,<br />

ed. D. T. Wicklow & B. Söderström. Berlin: Springer-<br />

Verlag, pp. 203 18.<br />

Kerry, B. R., Crump, D. H. & Mullen, L. A. (1982). Studies<br />

of the cereal cyst-nema<strong>to</strong>de, Heterodera avenae under<br />

continuous cereals, 1975 1978. II. Fungal parasitism<br />

of nema<strong>to</strong>de females and eggs. Annals of Applied<br />

Biology, 100, 489 499.<br />

Kerwin, J. L. & Washino, R. K. (1986).<br />

Oosporogenesis by Lagenidium giganteum: induction<br />

and maturation are regulated by calcium and<br />

calmodulin. Canadian Journal of Microbiology, 32,<br />

663 672.<br />

Kerwin, J. L., Johnson, L. M., Whisler, H. C. &<br />

Tuininga, A. R. (1992). Infection and morphogenesis<br />

of Pythium marinum in species of Porphyra and<br />

other red algae. Canadian Journal of Botany, 70,<br />

1017 1024.<br />

Keskin, B. & Fuchs, W. H. (1969). Der Infektionsvorgang<br />

bei Polymyxa betae. Archiv für Mikrobiologie, 68,<br />

218 226.<br />

Kessin, R. H. (2001). Dictyostelium. Evolution, Cell Biology,<br />

and the Development of Multicellularity. Cambridge:<br />

Cambridge University Press.<br />

Khairi, S. M. & Preece, T. F. (1978). Hawthorn powdery<br />

mildew: overwintering mycelium in buds and the<br />

effect of clipping hedges on disease epidemiology.<br />

Transactions of the British Mycological Society, 71,<br />

399 404.<br />

Khan, S. R. & Kimbrough, J. W. (1982). A reevaluation of<br />

the basidiomycetes based upon septal and basidial<br />

structures. Mycotaxon, 15, 103 120.<br />

Khan, S. R. & Talbot, P. H. B. (1975). Monosporous<br />

sporangiola in Mycotypha and<br />

Cunninghamella. Transactions of the British<br />

Mycological Society, 65, 29 39.<br />

Kher, K., Greening, J. P., Hat<strong>to</strong>n, J. P., Frazer, L. N. &<br />

Moore, D. (1992). Kinetics and mechanism of stem<br />

gravitropism in Coprinus cinereus. Mycological Research,<br />

96, 817 824.<br />

Kibby, G. & Bingham, J. (2004). Fungal portraits. No. 20:<br />

Phallus impudicus var. <strong>to</strong>gatus. Field Mycology, 5,<br />

111 112.<br />

Kieslich, K. (1997). Biotransformations. In Fungal<br />

Biotechnology, ed. T. Anke. London: Chapman & Hall,<br />

pp. 297 399.<br />

Kile, G. A. (1993). Plant diseases caused by species of<br />

Cera<strong>to</strong>cystis sensu stric<strong>to</strong> and Chalara. InCera<strong>to</strong>cystis and<br />

Ophios<strong>to</strong>ma: Taxonomy, Ecology and Pathology, ed. M. J.<br />

Wingfield, K. A. Seifert & J. F. Webber. St. Paul: APS<br />

Press, pp. 173 83.<br />

Killian, C. (1919). Sur la sexualité de l’ergot de Seigle, le<br />

Claviceps purpurea (Tulasne). Bulletin de la Société<br />

Mycologique de France, 35, 182 197.<br />

Kim, J. R., Yeon, S. H. & Ahny, J. (2002). Larvicidal<br />

activity against Plutella xylostella of cordycepin from<br />

the fruiting body of Cordyceps militaris. Pest<br />

Management Science, 58, 713 717.


756 REFERENCES<br />

Kimbeng, C. A. (1999). Genetic basis of crown rust<br />

resistance in perennial ryegrass, breeding strategies,<br />

and genetic variation among pathogen populations:<br />

a review. Australian Journal of Experimental Agriculture,<br />

39, 361 378.<br />

Kimbrough, J. W. (1994). Septal ultrastructure and<br />

ascomycete systematics. In Ascomycete Systematics.<br />

Problems and Perspectives in the Nineties, ed. D. L.<br />

Hawksworth. New York: Plenum Press, pp. 127 141.<br />

King, D. S. (1977). Systematics of Conidiobolus<br />

(En<strong>to</strong>mophthorales) using numerical taxonomy. III.<br />

Descriptions of recognized species. Canadian Journal<br />

of Botany, 55, 718 729.<br />

King, J. E. & Coley-Smith, J. R. (1969). Production of<br />

volatile alkyl sulphides by microbial degradation of<br />

synthetic alliin and alliin-like compounds, in relation<br />

<strong>to</strong> germination of sclerotia of Sclerotium cepivorum<br />

Berk. Annals of Applied Biology, 64, 303 314.<br />

Kinloch, B. B. (2003). White pine blister rust in North<br />

America: past and prognosis. Phy<strong>to</strong>pathology, 93,<br />

1044 1047.<br />

Kinloch, B. B. & Dupper, G. E. (2002). Genetic specificity<br />

in the white pine-blister rust pathosystem.<br />

Phy<strong>to</strong>pathology, 92, 278 280.<br />

Kinugawa, K. (1965). On the growth of Dictyophora<br />

indusiata. II. Relations between the change in<br />

osmotic value of expressed sap and the conversion of<br />

glycogen <strong>to</strong> reducing sugar in tissues during<br />

receptaculum elongation. Botanical Magazine (Tokyo),<br />

78, 240 244.<br />

Kirby, E. J. M. (1961). Host parasite relations in the<br />

choke disease of grasses. Transactions of the British<br />

Mycological Society, 44, 493 503.<br />

Kirby, J. J. H., Webster, J. & Baker, J. H. (1990). A particle<br />

plating method for the analysis of fungal community<br />

composition and structure. Mycological Research,<br />

94, 621 626.<br />

Kirk, P. M. (1977). Scanning electron microscopy of<br />

zygospore formation in Choanephora circinans<br />

(Mucorales). Transactions of the British Mycological<br />

Society, 68, 429 434.<br />

Kirk, P. M. (1984). A monograph of the<br />

Choanephoraceae. Mycological Papers, 152, 1 61. Kew:<br />

Commonwealth Mycological Institute.<br />

Kirk, P. M. (1993). Distribution of Zygomycetes<br />

the tropical connection. In Aspects of Tropical Mycology,<br />

ed. S. Isaac, J. C. Frankland, R. Watling & A. J. S.<br />

Whalley. Cambridge: Cambridge University Press,<br />

pp. 91 102.<br />

Kirk, P. M., Cannon, P. F., David, J. C. & Stalpers, J. A.,<br />

eds. (2001). Dictionary of the <strong>Fungi</strong>, 9th edn.<br />

Wallingford, UK: CABI Publishing.<br />

Kirk, T. K. & Cullen, D. (1998). Enzymology and<br />

molecular genetics of wood degradation by white-rot<br />

fungi. In Environmentally Friendly Technologies for the<br />

Pulp and Paper Industry, ed. R. A. Young & M. Akhtar.<br />

New York: Wiley, pp. 273 307.<br />

Kislev, M. E. (1982). Stem rust of wheat 3300 years old<br />

found in Israel. Science, 216, 993 994.<br />

Kiss, L. (1997). Graminicolous powdery mildew fungi as<br />

new natural hosts of Ampelomyces parasites. Canadian<br />

Journal of Botany, 75, 680 683.<br />

Kitamo<strong>to</strong>, K., Yoshizawa, K., Ohsumi, Y. & Anraku, Y.<br />

(1988). Dynamic aspects of vacuolar and cy<strong>to</strong>solic<br />

amino acid pools of Saccharomyces cerevisiae. Journal of<br />

Bacteriology, 170, 2683 2686.<br />

Kitancharoen, N., Hatai, K., Ogihara, R. & Aye, D. N. N.<br />

(1995). A new record of Achlya klebsiana from snakehead,<br />

Channa striatus, with fungal infection in<br />

Myammar. Mycoscience, 36, 235 238.<br />

Klein, D. & Eveleigh, D. E. (1998). Ecology of Trichoderma.<br />

In Trichoderma and Gliocladium. I. Basic Biology,<br />

Taxonomy and Genetics, ed. C. P. Kubicek & G. E.<br />

Harman. London: Taylor & Francis, pp. 57 74.<br />

Klein, S., Sherman, A. & Simchen, G. (1994). Regulation<br />

of meiosis and sporulation in Saccharomyces cerevisiae.<br />

In The Mycota I: Growth, Differentiation and Sexuality, ed.<br />

J. G. H. Wessels & F. Meinhardt. Berlin: Springer-<br />

Verlag, pp. 235 50.<br />

Klich, M. A. (2002). Identification of Common Aspergillus<br />

Species. Utrecht: Centraalbureau voor<br />

Schimmelcultures.<br />

Klich, M. A. & Cleveland, T. E. (2000). Aspergillus<br />

systematics and the molecular genetics of myco<strong>to</strong>xin<br />

biosynthesis. In Integration of Modern Taxonomic<br />

Methods for Penicillium and Aspergillus Classification, ed.<br />

R. A. Samson & J. I. Pitt. Amsterdam: Harwood<br />

Academic Publishers, pp. 425 34.<br />

Klionsky, D. J. (1997). Protein transport from the<br />

cy<strong>to</strong>plasm in<strong>to</strong> the vacuole. Journal of Membrane<br />

Biology, 157, 105 115.<br />

Klionsky, D. J., Herman, P. K. & Emr, S. D. (1990). The<br />

fungal vacuole: composition, function, and biogenesis.<br />

Microbiological Reviews, 54, 266 292.<br />

Kluczewski, S. M. & Lucas, J. A. (1983). Host infection and<br />

oospore formation by Peronospora parasitica in agricultural<br />

and horticultural Brassica species. Transactions<br />

of the British Mycological Society, 81, 591 596.<br />

Kluyver, A. J. & van Niel, C. B. (1927). Sporobolomyces: ein<br />

Basidiomyzet? Annales Mycologici, 25, 389 394.<br />

Kniep, H. (1919). Untersuchungen über den<br />

Antherenbrand (Ustilago violacea Pers.). Ein Beitrag<br />

zum Sexualitätsproblem. Zeitschrift für Botanik, 11,<br />

275 284.


REFERENCES<br />

757<br />

Ko, W. H. (1980). Hormonal regulation of sexual<br />

reproduction in Phy<strong>to</strong>phthora. Journal of General<br />

Microbiology, 116, 459 463.<br />

Ko, W. H., Lee, C. J. & Su, H. J. (1986). Chemical<br />

regulation of mating type in Phy<strong>to</strong>phthora parasitica.<br />

Mycologia, 78, 134 136.<br />

Kobayasi, Y. (1982). Keys <strong>to</strong> the taxa of the genus<br />

Cordyceps and Torrubiella. Transactions of the Mycological<br />

Society of Japan, 23, 329 364.<br />

Koch, E. & Slusarenko, A. (1990). Arabidopsis is susceptible<br />

<strong>to</strong> infection by a downy mildew fungus. Plant<br />

Cell, 2, 437 445.<br />

Köhler, E. (1956). Zur Kenntniss der Sexualität bei<br />

Synchytrium. Berichte der Deutschen Botanischen<br />

Gesellschaft, 69, 121 127.<br />

Kohlwein, S. D. (2000). The beauty of the yeast: live cell<br />

microscopy at the limits of optical resolution.<br />

Microscopy Research and Technique, 51, 511 529.<br />

Kohn, L. M. (1979). A monographic revision of the genus<br />

Sclerotinia. Mycotaxon, 9, 365 444.<br />

Kohn, L. M. (1992). Developing new characters for<br />

fungal systematics: an experimental approach for<br />

determining the rank of resolution. Mycologia, 84,<br />

139 163.<br />

Kohn, L. M. & Grenville, D. J. (1989). Ana<strong>to</strong>my and<br />

his<strong>to</strong>chemistry of stromatal anamorphs in the<br />

Sclerotiniaceae. Canadian Journal of Botany, 67,<br />

371 393.<br />

Kole, A. P. (1965). Resting-spore germination in<br />

Synchytrium endobioticum. Netherlands Journal of Plant<br />

Pathology, 71, 72 78.<br />

Kollar, A. (1998). A simple method <strong>to</strong> forecast the<br />

ascospore discharge of Venturia inaequalis. Zeitschrift<br />

für Pflanzenkrankheiten und Pflanzenschutz, 105,<br />

489 495.<br />

Kollár, R., Reinhold, B. B., Petráková, E., et al. (1997).<br />

Architecture of the yeast cell wall: b-(1,6)-glucan<br />

interconnects mannoprotein, b-(1,3)-glucan,<br />

and chitin. Journal of Biological Chemistry, 272,<br />

17 762 775.<br />

Kolmer, J. A. (2005). Tracking wheat rust on a continental<br />

scale. Current Opinion in Plant Biology, 8,<br />

441 449.<br />

Koltin, Y. & Day, P. R. (1975). Specificity of Ustilago<br />

maydis killer proteins. Applied Microbiology, 30,<br />

694 696.<br />

Kombrink, E. & Somssich, I. E. (1997). Pathogenesisrelated<br />

proteins and plant defense. In The Mycota VA:<br />

Plant Relationships, ed. G. C. Carroll & P. Tudzynski.<br />

Berlin: Springer-Verlag, pp. 107 28.<br />

Konijn, T. M., van der Meene, J. G. C., Bonner, J. T. &<br />

Barkley, D. S. (1967). The acrasin activity of<br />

adenosine-3’-5’ cyclic phosphate. Proceedings of the<br />

National Academy of Sciences of the USA, 58, 1152 1154.<br />

Konishi, M. (1967). Growth-promoting effects of certain<br />

amino acids on the Agaricus fruit body. Mushroom<br />

Science, 6, 121 134.<br />

Kono, Y., Sekido, S., Yamaguchi, I., et al. (1991).<br />

Structures of two novel pyriculol-related compounds<br />

and identification of naturally produced epipyriculol<br />

from Pyricularia oryzae. Agricultural and Biological<br />

Chemistry, 55, 2785 2791.<br />

Kope, H. H. & Fortin, J. A. (1990). Germination and<br />

comparative morphology of basidiospores of<br />

Pisolithus arhizus. Mycologia, 82, 350 357.<br />

Kopecka, M., Fleet, G. H. & Phaff, H. J. (1995).<br />

Ultrastructure of the cell wall of Schizosaccharomyces<br />

pombe following treatment with various glucanases.<br />

Journal of Structural Biology, 114, 140 152.<br />

Korf, R. P. (1972). Synoptic key <strong>to</strong> genera of the<br />

Pezizales. Mycologia, 64, 937 993.<br />

Korhonen, K. & Hintikka, V. (1974). Cy<strong>to</strong>logical<br />

evidence for somatic diploidization in dikaryotic<br />

cells of Armillaria mellea. Archiv für Mikrobiologie, 95,<br />

187 192.<br />

Korhonen, K. & Stenlid, J. (1998). Biology of<br />

Heterobasidion annosum. In Heterobasidion annosum.<br />

Biology, Ecology, Impact and Control, ed. S. Woodward, J.<br />

Stenlid, R. Karjalainan & A. Hüttermann.<br />

Wallingford, UK: CAB International, pp. 43 70.<br />

Koske, R. E. & Duncan, I. W. (1974). Temperature effects<br />

on growth, sporulation and germination of some<br />

aquatic hyphomycetes. Canadian Journal of Botany, 52,<br />

1387 1391.<br />

Kothe, E. (1996). Tetrapolar fungal mating types: sexes<br />

by the thousands. FEMS Microbiology Reviews, 18,<br />

65 87.<br />

Kothe, E. (1999). Mating types and pheromone recognition<br />

in the homobasidiomycete Schizophyllum<br />

commune. Fungal Genetics and Biology, 27, 146 152.<br />

Kramer, C. L. (1961). Morphological development and<br />

nuclear behaviour in the genus Taphrina. Mycologia,<br />

52, 295 320.<br />

Kramer, C. L. (1982). Production, release and dispersal<br />

of basidiospores. In Decomposer Basidiomycetes, their<br />

Biology and Ecology, ed. J. C. Frankland, J. N. Hedger &<br />

M. J. Swift. Cambridge: Cambridge University Press,<br />

pp. 33 49.<br />

Kramer, C. L. (1987). The Taphrinales. Studies in Mycology,<br />

30, 151 166.<br />

Krasnoff, S. B. & Gupta, S. (1992). Efrapeptin production<br />

by Tolypocladium fungi (Deuteromycotina,<br />

Hyphomycetes) intraspecific and interspecific<br />

variation. Journal of Chemical Ecology, 18, 1727 1741.


758 REFERENCES<br />

Krassilov, V. A. & Makulbekov, N. M. (2003). The first<br />

finding of Gasteromycetes in the Cretaceous of<br />

Mongolia. Palaeon<strong>to</strong>logical Journal, 37, 439 442.<br />

Kreger-van Rij, N. J. W. & Veenhuis, M. (1971). A<br />

comparative study of the cell wall structure of<br />

basidiomyce<strong>to</strong>us and related yeasts. Journal of General<br />

Microbiology, 68, 87 95.<br />

Kreger-van Rij, N. J.W., Veenhuis, M. & Leemburg-van<br />

der Graaf, C. A. (1974). Ultrastructure of hyphae and<br />

ascospores in the genus Eremascus Eidam. An<strong>to</strong>nie van<br />

Leeuwenhoek, 40, 533 542.<br />

Krejzová, R. (1978). Taxonomy, morphology and surface<br />

structure of Basidiobolus sp. isolate. Journal of<br />

Invertebrate Pathology, 31, 157 163.<br />

Křen, V. & Cvak, L., eds. (1999). Ergot. The Genus Claviceps.<br />

Amsterdam: Harwood Academic Publishers.<br />

Kronstad, J. W. & Staben, C. (1997). Mating-type in<br />

filamen<strong>to</strong>us fungi. Annual Review of Genetics, 31,<br />

245 276.<br />

Kronstad, J., de Maria, A. D., Funnell, D., et al. (1998).<br />

Signaling via cAMP in fungi: interconnections with<br />

mi<strong>to</strong>gen-activated protein kinase pathways. Archives<br />

of Microbiology, 170, 395 404.<br />

Kroon, L. P. N. M., Bakker, F. T., van den Bosch, G. B. M.,<br />

Bonants, P. J. M. & Flier, W. G. (2004). Phylogenetic<br />

analysis of Phy<strong>to</strong>phthora species based on mi<strong>to</strong>chondrial<br />

and nuclear DNA sequences. Fungal Genetics and<br />

Biology, 41, 766 782.<br />

Kropf, D. L. & Harold, F. M. (1982). Selective transport of<br />

nutrients via rhizoids of the water mold<br />

Blas<strong>to</strong>cladiella emersonii. Journal of Bacteriology, 151,<br />

429 437.<br />

Kropp, B. R. & Mueller, G. M. (1999). Laccaria. In<br />

Ec<strong>to</strong>mycorrhizal <strong>Fungi</strong>. Key Genera in Profile, ed. J. W. G.<br />

Cairney & S. M. Chambers. Berlin: Springer-Verlag,<br />

pp. 65 88.<br />

Krüger, D., Binder, M., Fischer, M. & Kreisel, H. (2001).<br />

The Lycoperdales. A molecular approach <strong>to</strong> the<br />

systematics of some gasteroid mushrooms. Mycologia,<br />

93, 947 957.<br />

Kubicek, C. P., Eveleigh, D. E., Esterbauer, H., Steiner,<br />

W. & Kubicek-Pranz, E. M., eds. (1990). Trichoderma<br />

reesii: Cellulases, Biodiversity, Genetics, Physiology and<br />

Applications. Cambridge: Royal Society of Chemistry.<br />

Kuck, K. H., Scheinpflug, H. & Pontzen, R. (1995). DMI<br />

fungicides. In Modern Selective <strong>Fungi</strong>cides. Properties,<br />

Applications, Mechanisms of Action, 2nd edn, ed. H. Lyr.<br />

New York: Gustav Fischer Verlag, pp. 205 258.<br />

Kucsera, J., Pfeiffer, I. & Ferenczy, L. (1998). Homothallic<br />

life cycle in the diploid red yeast Xanthophyllomyces<br />

dendrorhous (Phaffia rhodozyma). An<strong>to</strong>nie van<br />

Leeuwenhoek, 73, 163 168.<br />

Kucsera, J., Pfeiffer, I. & Takeo, K. (2000). Biology of the<br />

red yeast Xanthophyllomyces dendrorhous (Phaffia<br />

rhodozyma). Mycoscience, 41, 195 199.<br />

Kuehn, H. H. (1956). Observations on the<br />

Gymnoascaceae. III. Developmental morphology of<br />

Gymnoascus reessii, a new species of Gymnoascus and<br />

Eidamella deflexa. Mycologia, 48, 805 820.<br />

Kües, U. (2000). Life his<strong>to</strong>ry and developmental<br />

processes in the basidiomycete Coprinus cinereus.<br />

Microbiology and Molecular Biology Reviews, 64,<br />

316 353.<br />

Kües, U. & Cassel<strong>to</strong>n, L. A. (1992). Fungal mating-type<br />

genes regula<strong>to</strong>rs of sexual development.<br />

Mycological Research, 96, 993 1006.<br />

Kuhlman, E. G. (1972). Variation in zygospore formation<br />

among species of Mortierella. Mycologia, 64,<br />

325 341.<br />

Kuhls, K., Lieckfeldt, E., Samuels, G. J., et al. (1996).<br />

Molecular evidence that the asexual industrial<br />

fungus Trichoderma reesei is a clonal derivative of the<br />

ascomycete Hypocrea jecorina. Proceedings of the<br />

National Academy of Sciences of the USA, 93, 7755 7760.<br />

Kuldau, G. & Bacon, C. W. (2001). Claviceps and related<br />

fungi. In Foodborne Disease Handbook 3: Plant Toxicants,<br />

2nd edn, ed. H. Hui, R. A. Smith & D. G. Spoerke. New<br />

York: Marcel Dekker Inc., pp. 503 534.<br />

Kuldau, G. A., Liu, J., White, F., Siegel, M. R. & Schardl,<br />

C. L. (1997). Molecular systematics of Clavicipitaceae<br />

supporting monophyly of genus Epichloe and form<br />

genus Ephelis. Mycologia, 89, 431 441.<br />

Kulka, M. & von Schmeling, B. (1995). Carboxin<br />

fungicides and related compounds. In Modern<br />

Selective <strong>Fungi</strong>cides. Properties, Applications, Mechanisms<br />

of Action, ed. H. Lyr. Jena: Gustav Fischer Verlag,<br />

pp. 133 147.<br />

Kulkarni, U. K. (1963). Initiation of the dikaryon in<br />

Claviceps microcephala (Wallr.) Tul. Mycopathologia et<br />

Mycologia Applicata, 21, 19 22.<br />

Kullnig-Gradinger, C. M., Szakags, G. & Kubicek, C. P.<br />

(2002). Phylogeny and evolution of the genus<br />

Trichoderma: a multigene approach. Mycological<br />

Research, 106, 757 767.<br />

Kumar, S. & Rzhetsky, A. (1996). Evolutionary relationships<br />

of eukaryotic kingdoms. Journal of Molecular<br />

Evolution, 42, 183 193.<br />

Kumar, J., Hückelhoven, R., Beckhove, U., Nagarajan, S.<br />

& Kogel, K.-H. (2001). A compromised Mlo pathway<br />

affects the response of barley <strong>to</strong> the necrotrophic<br />

fungus Bipolaris sorokiniana (teleomorph: Cochliobolus<br />

sativus) and its <strong>to</strong>xins. Phy<strong>to</strong>pathology, 91, 127 133.<br />

Kumar, J., Schäfer, P., Hückelhoven, R., et al. (2002).<br />

Bipolaris sorokiniana, a cereal pathogen of global


REFERENCES<br />

759<br />

concern: cy<strong>to</strong>logical and molecular approaches<br />

<strong>to</strong>wards better control. Molecular Plant Pathology, 3,<br />

185 195.<br />

Kunert, J. (2000). Physiology of keratinolytic fungi. In<br />

Biology of Derma<strong>to</strong>phytes and Other Keratinolytic <strong>Fungi</strong>,<br />

ed. R. K. S. Kushwara & J. Guarro. Bilbao: Revista<br />

Iberoamericana de Micología, pp. 77 85.<br />

Kuninaga, S., Natsuaki, T., Takeuchi, T. & Yokosawa, R.<br />

(1997). Sequence variation of the rDNA ITS regions<br />

within and between anas<strong>to</strong>mosis groups in<br />

Rhizoc<strong>to</strong>nia solani. Current Genetics, 32, 237 243.<br />

Kunoh, H. & Ishizaki, H. (1981). Cy<strong>to</strong>logical studies of<br />

early stages of powdery mildew in barley and wheat.<br />

VII. Reciprocal translocation of a fluorescent dye<br />

between barley coleoptile cells and conidia.<br />

Physiological Plant Pathology, 18, 207 211.<br />

Kurahashi, H., Imai, Y. & Yamamo<strong>to</strong>, M. (2002).<br />

Tropomyosin is required for the cell fusion process<br />

during conjugation in fission yeast. Genes <strong>to</strong> Cells, 7,<br />

375 384.<br />

Kurtzman, C. P. (1995). Relationships among the genera<br />

Ashbya, Eremothecium, Holleya and Nema<strong>to</strong>spora determined<br />

from rDNA sequence divergence. Journal of<br />

Industrial Microbiology, 14, 523 530.<br />

Kurtzman, C. P. (1998). Pichia E. C. Hansen emend.<br />

Kurtzman. In The Yeasts: A Taxonomic Study, 4th edn,<br />

ed. C. P. Kurtzman & J. W. Fell. Amsterdam: Elsevier,<br />

pp. 273 352.<br />

Kurtzman, C. P. & Fell, J. W., eds. (1998). The Yeasts: A<br />

Taxonomic Study, 4th edn. Amsterdam: Elsevier.<br />

Kurtzman, C. P. & Robnett, C. J. (1998). Identification<br />

and phylogeny of ascomyce<strong>to</strong>us yeasts from analysis<br />

of nuclear large subunit (26S) ribosomal DNA<br />

partial sequences. An<strong>to</strong>nie van Leeuwenhoek, 73,<br />

331 371.<br />

Kurtzman, C. P. & Robnett, C. J. (2003). Phylogenetic<br />

relationships among yeasts of the ‘Saccharomyces<br />

complex’ determined from multigene sequence<br />

analyses. FEMS Yeast Research, 3, 417 432.<br />

Kurtzman, C. P. & Smith, M. T. (1998). Saccharomycopsis<br />

Schöning. In The Yeasts: A Taxonomic Study, 4th edn,<br />

ed. C. P. Kurtzman & J. W. Fell. Amsterdam: Elsevier,<br />

pp. 374 86.<br />

Kurtzman, C. P. & Sugiyama, J. (2001). Ascomyce<strong>to</strong>us<br />

yeasts and yeastlike taxa. In The Mycota VIIA:<br />

Systematics and Evolution, ed. D. J. McLaughlin, E. G.<br />

McLaughlin & P. A. Lemke. Berlin: Springer-Verlag,<br />

pp. 179 200.<br />

Kusch, J., Meyer, A., Snyder, M. P. & Barral, Y. (2002).<br />

Microtubule capture by the cleavage apparatus is<br />

required for proper spindle positioning in yeast.<br />

Genes and Development, 16, 1627 1639.<br />

Kushalappa, A. H. & Eskes, A. B. (1989). Advances in<br />

coffee rust research. Annual Review of Phy<strong>to</strong>pathology,<br />

27, 503 531.<br />

Kwasna, H. & Kosiak, B. (2003). Lewia avenicola sp. nov.,<br />

and its Alternaria anamorph from oat grain with a<br />

key <strong>to</strong> the species of Lewia. Mycological Research, 107,<br />

371 376.<br />

Kwok, O. C. H., Plattner, R., Weisleder, D. & Wicklow,<br />

D. T. (1992). A nematicidal <strong>to</strong>xin from Pleurotus<br />

ostreatus NRRL 3526. Journal of Chemical Ecology, 18,<br />

127 136.<br />

Kwon-Chung, K. J. (1973). Studies on Emmonsiella capsulata.<br />

I. Heterothallism and development of the<br />

ascocarp. Mycologia, 65, 109 121.<br />

Kwon-Chung, K. J. (1998). Filobasidiella Kwon-Chung. In<br />

The Yeasts: A Taxonomic Study, 4th edn, ed. C. P.<br />

Kurtzman & J. W. Fell. Amsterdam: Elsevier,<br />

pp. 656 662.<br />

Kwon-Chung, K. J. & Bennett, J. E. (1978). Distribution of<br />

a and a mating types of Cryp<strong>to</strong>coccus neoformans<br />

among natural and clinical isolates. American Journal<br />

of Epidemiology, 108, 337 340.<br />

Kwon-Chung, K. J. & Bennett, J. E. (1992). Medical<br />

Mycology. Malvern, USA: Lea and Febiger.<br />

Kwon-Chung, K. J., Edman, J. C. & Wickes, B. L. (1992).<br />

Genetic association of mating types and virulence in<br />

Cryp<strong>to</strong>coccus neoformans. Infection and Immunity, 60,<br />

602 605.<br />

Kybal, J. (1964). Changes in N and P content during<br />

growth of ergot sclerotia (Claviceps purpurea) due <strong>to</strong><br />

nutrition supplied by rye. Phy<strong>to</strong>pathology, 54, 244 245.<br />

Labeyrie, E. S., Molloy, D. P. & Lichtwardt, R. W. (1996).<br />

An investigation of Harpellales (Trichomycetes) in<br />

New York State blackflies (Diptera: Simuliidae).<br />

Journal of Invertebrate Pathology, 68, 293 298.<br />

Lacey, J. (1994). Aspergilli in feeds and seeds. In The<br />

Genus Aspergillus. From Taxonomy and Genetics <strong>to</strong><br />

Industrial Application, ed. K. A. Powell, A. Renwick &<br />

J. F. Peberdy. New York: Plenum Press, pp. 73 92.<br />

Lacey, J. (1996). Spore dispersal its ecology and<br />

disease: the British contribution <strong>to</strong> fungal aerobiology.<br />

Mycological Research, 100, 641 660.<br />

Lachance, M.-A., Pupovac-Velikonja, A., Natajan, S. &<br />

Schlag-Edler, B. (2000). Nutrition and phylogeny of<br />

predacious yeasts. Canadian Journal of Microbiology, 46,<br />

495 505.<br />

Lachke, S. A., Lockhart, S. R., Daniels, K. J. & Soll, D. R.<br />

(2003). Skin facilitates Candida albicans mating.<br />

Infection and Immunity, 71, 4970 4976.<br />

Lafon, R. & Bulit, J. (1981). Downy mildew of the vine. In<br />

The Downy Mildews, ed. D. M. Spencer. London:<br />

Academic Press, pp. 601 614.


760 REFERENCES<br />

Laidlaw, W. M. R. (1985). A method for the detection of<br />

the resting sporangia of pota<strong>to</strong> wart disease<br />

(Synchytrium endobioticum) in the soil of old outbreak<br />

sites. Pota<strong>to</strong> Research, 28, 223 232.<br />

Laing, S. A. K. & Deacon, J. W. (1991). Video microscopical<br />

comparison of mycoparasitism by Pythium<br />

oligandrum, P. nunn and an unnamed Pythium species.<br />

Mycological Research, 95, 469 479.<br />

Lamb, B. C. (1996). Ascomycete genetics: the part played<br />

by ascus segregation phenomena in our understanding<br />

of the mechanisms of recombination.<br />

Mycological Research, 100, 1025 1059.<br />

Lampila, L. E., Wallen, S. E. & Bullermann, L. B. (1985). A<br />

review of fac<strong>to</strong>rs affecting biosynthesis of carotenoids<br />

by the order Mucorales. Mycopathologia, 90,<br />

65 80.<br />

Landvik, S., Egger, K. N. & Schumacher, T. (1997).<br />

Towards a subordinal classification of the Pezizales<br />

(Ascomycota): phylogenetic analyses of SSU rDNA<br />

sequences. Nordic Journal of Botany, 17, 403 418.<br />

Landvik, S., Schumacher, T. K., Eriksson, O. E. & Moss,<br />

S. T. (2003). Morphology and ultrastructure of<br />

Neolecta species. Mycological Research, 107, 1021 1031.<br />

Langan, T. A., Gautier, J., Lohka, M., et al. (1989).<br />

Mammalian growth-associated H1 his<strong>to</strong>ne kinase: a<br />

homolog of cdc2 þ /CDC28 protein kinases controlling<br />

mi<strong>to</strong>tic entry in yeast and frog cells. Molecular and<br />

Cellular Biology, 9, 3860 3868.<br />

Lange, L. (1987). Synchytrium endobioticum. In Zoosporic<br />

<strong>Fungi</strong> in Teaching and Research, ed. M. S. Fuller & A.<br />

Jaworski. Athens, USA: Southeastern Publishing<br />

Corporation, pp. 24 5.<br />

Lange, L. & Insunza, V. (1977). Root-inhabiting Olpidium<br />

species: the O. radicale complex. Transactions of the<br />

British Mycological Society, 69, 377 384.<br />

Lange, L. & Olson, L. W. (1976a). The flagellar apparatus<br />

and striated rhizoplast of the zoospore of Olpidium<br />

brassicae. Pro<strong>to</strong>plasma, 89, 339 351.<br />

Lange, L. & Olson, L. W. (1976b). The zoospore of<br />

Olpidium brassicae. Pro<strong>to</strong>plasma, 90, 33 45.<br />

Lange, L. & Olson, L. W. (1978). The zoospore of<br />

Synchytrium endobioticum. Canadian Journal of Botany,<br />

56, 1229 1239.<br />

Lange, L. & Olson, L. W. (1979). The uniflagellate<br />

Phycomycete zoospore. Dansk Botanisk Arkiv, 33,<br />

7 95.<br />

Lange, L. & Olson, L. W. (1980). Transfer of the<br />

Physodermataceae from the Chytridiales <strong>to</strong> the<br />

Blas<strong>to</strong>cladiales. Transactions of the British Mycological<br />

Society, 74, 449 457.<br />

Lange, L. & Olson, L. W. (1981). Development of the<br />

resting sporangia of Synchytrium endobioticum, the<br />

causal agent of pota<strong>to</strong> wart disease. Pro<strong>to</strong>plasma, 106,<br />

83 95.<br />

Lange, L. & Olson, L. W. (1983). The fungal zoospore. Its<br />

structure and biological significance. In Zoosporic<br />

Plant Pathogens. A Modern Perspective, ed. S. T. Buczacki.<br />

London: Academic Press, pp. 1 42.<br />

Lange, L., Sparrow, F. K. & Olson, L. W. (1987).<br />

Physoderma spp. In Zoosporic <strong>Fungi</strong> in Teaching and<br />

Research, ed. M. S. Fuller & A. Jaworski. Athens, USA:<br />

Southeastern Publishing Corporation, pp. 36 7.<br />

Lange, M. (1952). Species concept in the genus Coprinus.<br />

Dansk Botanisk Arkiv, 14, 1 140.<br />

Lapeyrie, F. & Mendgen, K. (1993). Quantitative estimation<br />

of surface carbohydrates of ec<strong>to</strong>mycorrhizal<br />

fungi in pure culture and during Eucalyptus root<br />

infection. Mycological Research, 97, 603 609.<br />

Large, E. C. (1940). The Advance of the <strong>Fungi</strong>. London:<br />

Jonathan Cape.<br />

Larkin, R. P. & Fravel, D. R. (1999). Mechanisms of action<br />

and dose response relationships governing biological<br />

control of fusarium wilt of <strong>to</strong>ma<strong>to</strong> by<br />

nonpathogenic Fusarium spp. Phy<strong>to</strong>pathology, 89,<br />

1152 1161.<br />

Larsen, J., Mansfield-Giese, K. & Bødker, L. (2000).<br />

Quantification of Aphanomyces euteiches in pea roots<br />

using specific fatty acids. Mycological Research, 104,<br />

858 864.<br />

Larsen, M. (2000). Prospects for controlling animal<br />

parasitic nema<strong>to</strong>des by predacious micro fungi.<br />

Parasi<strong>to</strong>logy, 120, S121 S131.<br />

Larsson, E. & Larsson, K.-H. (2003). Phylogenetic<br />

relationships of russuloid basidiomycetes with<br />

emphasis on aphyllophoralean taxa. Mycologia, 95,<br />

1037 1065.<br />

Last, F. T., Mason, P. A., Wilson, J. & Deacon, J. W. (1983).<br />

Fine roots and sheathing mycorrhizas: their<br />

formation, function and dynamics. Plant and Soil,<br />

71, 9 21.<br />

Last, F. T., Digh<strong>to</strong>n, J. & Mason, P. A. (1987). Successions<br />

of sheathing mycorrhizal fungi. Trends in Ecology and<br />

Evolution, 2, 157 161.<br />

Latgé, J.-P. (1999). Aspergillus fumigatus and aspergillosis.<br />

Clinical Microbiology Reviews, 12, 310 350.<br />

Latgé, J.-P. (2001). The pathobiology of Aspergillus<br />

fumigatus. Trends in Microbiology, 9, 382 389.<br />

Latgé, J. P., Perry, D. F., Prévost, M. C. & Samson, R.<br />

(1989). Ultrastructural studies of primary spores of<br />

Conidiobolus, Erynia, and related En<strong>to</strong>mophthorales.<br />

Canadian Journal of Botany, 67, 2576 2589.<br />

Latunde-Dada, A. O. (2001). Colle<strong>to</strong>trichum: tales of<br />

forcible entry, stealth, transient confinement and<br />

breakout. Molecular Plant Pathology, 2, 187 198.


REFERENCES<br />

761<br />

Latunde-Dada, A. O., O’Connell, R. J., Nash, C., Pring,<br />

R. J., Lucas, J. A. & Bailey, J. A. (1996). Infection process<br />

and identity of the hemibiotrophic anthracnose<br />

fungus (Colle<strong>to</strong>trichum destructivum) from cowpea<br />

(Vigna unguiculata). Mycological Research, 100,<br />

1133 1141.<br />

Lawrey, J. D. (1986). Biological role of lichen substances.<br />

Bryologist, 89, 111 122.<br />

Lawrey, J. D. & Diederich, P. (2003). Lichenicolous fungi:<br />

interactions, evolution, and biodiversity. Bryologist,<br />

106, 80 120.<br />

Law<strong>to</strong>n, K., Weymann, K., Friedrich, L., Vernooij, B.,<br />

Uknes, S. & Ryals, J. (1995). Systemic acquired<br />

resistance in Arabidopsis requires salicylic acid but<br />

not ethylene. Molecular Plant Microbe Interactions, 8,<br />

863 870.<br />

Lazéra, M. S., Pires, F. D. A., Camillo-Coura, L., et al.<br />

(1996). Natural habitat of Cryp<strong>to</strong>coccus neoformans var.<br />

neoformans in decaying wood forming hollows in<br />

living trees. Journal of Medical and Veterinary Mycology,<br />

34, 127 131.<br />

Leach, C. M. (1968). An action spectrum for light<br />

inhibition of the ‘terminal phase’ of pho<strong>to</strong>sporogenesis<br />

in the fungus Stemphylium botryosum.<br />

Mycologia, 60, 532 546.<br />

Leach, C. M. (1971). Regulation of perithecium development<br />

in Pleospora herbarum by light and temperature.<br />

Transactions of the British Mycological Society, 57,<br />

295 315.<br />

Leach, C. M. (1976). An electrostatic theory <strong>to</strong> explain<br />

violent spore liberation by Drechslera turcica and<br />

other fungi. Mycologia, 68, 63 86.<br />

Leadbeater, G. & Mercer, C. (1957). Zygospores in<br />

Pip<strong>to</strong>cephalis. Transactions of the British Mycological<br />

Society, 40, 109 116.<br />

Leake, J., Johnson, D., et al. (2004). Networks of power<br />

and influence: the role of mycorrhizal mycelium in<br />

controlling plant communities and agroecosystem<br />

functioning. Canadian Journal of Botany, 82,<br />

1016 1045.<br />

Leander, C. A. & Porter, D. (2001). The<br />

Labyrinthulomycota is comprised of three distinct<br />

lineages. Mycologia, 93, 459 464.<br />

Leathers, T. D. (2003). Biotechnological production and<br />

applications of pullulan. Applied Microbiology and<br />

Biotechnology, 62, 468 473.<br />

Leberer, E., Thomas, D. Y. & Whiteway, M. (1997).<br />

Pheromone signalling and polarized morphogenesis<br />

in yeast. Current Opinion in Genetics and Development, 7,<br />

59 66.<br />

Leclerc, M. C., Guillot, G. & Deville, M. (2000).<br />

Taxonomic and phylogenetic analysis of<br />

Saprolegniaceae (Oomycetes) inferred from LSU<br />

rDNA and ITS sequence comparisons. An<strong>to</strong>nie van<br />

Leeuwenhoek, 77, 369 377.<br />

Lee, M. G. & Nurse, P. (1987). Complementation used <strong>to</strong><br />

clone a human homologue of the fission yeast cell<br />

cycle control gene cdc2. Nature, 327, 31 35.<br />

Lee, N., Bakkeren, G., Wong, K., Sherwood, J. E. &<br />

Kronstad, J. W. (1999). The mating-type and pathogenicity<br />

locus of the fungus Ustilago hordei spans a<br />

500-kb region. Proceedings of the National Academy of<br />

Sciences of the USA, 96, 15 026 15031.<br />

Legard, D. E., Lee, T. Y. & Fry, W. E. (1995). Pathogenic<br />

specialization in Phy<strong>to</strong>phthora infestans: aggressiveness<br />

on <strong>to</strong>ma<strong>to</strong>. Phy<strong>to</strong>pathology, 85, 1356 1361.<br />

Lehnen, L. P. & Powell, M. J. (1989). The role of<br />

kine<strong>to</strong>some-associated organelles in the attachment<br />

of encysting secondary zoospores of Saprolegnia ferax<br />

<strong>to</strong> substrates. Pro<strong>to</strong>plasma, 149, 163 174.<br />

Lei, J., Lapeyrie, F., Malajczuk, N. & Dexheimer, J. (1990).<br />

Infectivity of pine and eucalypt isolates of Pisolithus<br />

tinc<strong>to</strong>rius (Pers.) Coker & Couch on roots of Eucalyptus<br />

urophylla S. T. Blake in vitro. II. Ultrastructural and<br />

biochemical changes at the early stage of mycorrhiza<br />

formation. New Phy<strong>to</strong>logist, 116, 115 122.<br />

Leite, B. & Nicholson, R. L. (1992). Mycosporine-alanine:<br />

a self-inhibi<strong>to</strong>r of germination from the conidial<br />

mucilage of Colle<strong>to</strong>trichum graminicola. Experimental<br />

Mycology, 16, 76 86.<br />

Leith, I. D. & Fowler, D. (1988). Urban distribution of<br />

Rhytisma acerinum (Pers.) Fries (tar spot) on sycamore.<br />

New Phy<strong>to</strong>logist, 108, 175 181.<br />

LéJohn, H. B. (1972). Enzyme regulation, lysine pathways<br />

and cell wall structures as indica<strong>to</strong>rs of major<br />

lines of evolution in fungi. Nature, 231, 164 168.<br />

Lemke, A., Kiderlen, A. F. & Kayser, O. (2005).<br />

Amphotericin B. Applied Microbiology and<br />

Biotechnology, 68, 151 162.<br />

Lengeler, K. B., Fox, D. S., Fraser, J. A., et al. (2002).<br />

Mating-type locus of Cryp<strong>to</strong>coccus neoformans: a step in<br />

the evolution of sex chromosomes. Eukaryotic Cell, 1,<br />

704 718.<br />

LePage, B. A., Currah, R. S., S<strong>to</strong>ckney, R. A. & Rothwell,<br />

G. W. (1997). Fossil ec<strong>to</strong>mycorrhizae from the middle<br />

Eocene. American Journal of Botany, 84, 410 412.<br />

Leroux, P. & Gredt, M. (1997). Evolution of fungicide<br />

resistance in the cereal eyespot fungi Tapesia<br />

yallundae and T. acuformis in France. Pesticide Science,<br />

51, 321 327.<br />

Leroux, P., Fritz, R., Debieu, D., et al. (2002).<br />

Mechanisms of resistance <strong>to</strong> fungicides in field<br />

strains of Botrytis cinerea. Pest Management Science, 58,<br />

876 888.


762 REFERENCES<br />

Letscher-Bru, V. & Herbrecht, R. (2003). Caspofungin:<br />

the first representative of a new antifungal class.<br />

Journal of Antimicrobial Chemotherapy, 51, 513 521.<br />

Leuchtmann, A. (1984). Über Phaeosphaeria Miyake und<br />

andere bitunicate Ascomyceten mit mehrfach querseptierten<br />

Ascosporen. Sydowia, 37, 75 194.<br />

Leuchtmann, A. (2003). Taxonomy and diversity of<br />

Epichloe endophytes. In Clavicipitalean <strong>Fungi</strong>, ed.<br />

J. F. White, C. W. Bacon, N. L. Hywel-Jones &<br />

J. W. Spatafora. New York: Marcel Dekker Inc.,<br />

pp. 169 194.<br />

Leuchtmann, A., Schardl, C. L. & Siegel, M. R. (1994).<br />

Sexual compatibility and taxonomy of a new species<br />

of Epichloe symbiotic with fine fescue grasses.<br />

Mycologia, 86, 802 812.<br />

Léveillé, J. H. (1851). Organisation et disposition méthodique<br />

des espèces qui composent le genre Erysiphé.<br />

Annales des Sciences Naturelles. Botanique, Séries 3, 15,<br />

109 179.<br />

Lévesque, C. A. & de Cock, A. W. A. M. (2004). Molecular<br />

phylogeny and taxonomy of the genus Pythium.<br />

Mycological Research, 108, 1363 1383.<br />

Levetin, E. & Caroselli, N. E. (1976). A simplified<br />

medium for growth and sporulation of Pilobolus<br />

species. Mycologia, 68, 1254 1258.<br />

Levings, C. S., Rhoads, D. M. & Siedow, J. N. (1995).<br />

Molecular interactions of Bipolaris maydis T-<strong>to</strong>xin<br />

and maize. Canadian Journal of Botany, 73, S483 S489.<br />

Levisohn, I. (1927). Beitrag zur Entwicklungsgeschichte<br />

und Biologie von Basidiobolus ranarum Eidam.<br />

Jahrbücher für Wissenschaftliche Botanik, 66, 513 555.<br />

Lewin, B. (2000). Genes VII. New York: Oxford University<br />

Press.<br />

Lewis, L. A. & Decaris, B. (1974). The induction of<br />

apothecial formation in Ascobolus immersus by a<br />

spermatization technique. Transactions of the British<br />

Mycological Society, 63, 197 199.<br />

Lewis, T. E., Nichols, P. D. & McMeekin, T. A. (1999). The<br />

biotechnological potential of thraus<strong>to</strong>chytrids.<br />

Marine Biotechnology, 1, 580 587.<br />

Li, J., Heath, I. B. & Packer, L. (1993). The phylogenetic<br />

relationship of the anaerobic chytridiomyce<strong>to</strong>us gut<br />

fungi (Neocallimasticaceae) and the<br />

Chytridiomycota. II. Cladistic analysis of structural<br />

data and description of the Neocallimasticales ord.<br />

nov. Canadian Journal of Botany, 71, 393 407.<br />

Li, L., Gerecke, E. E. & Zolan, M. E. (1999). Homolog<br />

pairing and meiotic progression in Coprinus cinereus.<br />

Chromosoma, 108, 384 392.<br />

Li, N., Erman, M., Pangborn, W., et al. (1999). Structure<br />

of Ustilago maydis killer <strong>to</strong>xin KP6 a-subunit. Journal of<br />

Biological Chemistry, 274, 20 425 431.<br />

Lichtwardt, R. W. (1967). Zygospores and spore appendages<br />

of Harpella (Trichomycetes) from larvae of<br />

Simuliidae. Mycologia, 59, 482 491.<br />

Lichtwardt, R. W. (1986). The Trichomycetes. Fungal<br />

Associates of Arthropods. Berlin: Springer-Verlag.<br />

Lichtwardt, R. W. (1996). Trichomycetes and the<br />

arthropod gut. In The Mycota VI: Human and Animal<br />

Relationships, ed. D. J. Howard & J. D. Miller. Berlin:<br />

Springer-Verlag, pp. 315 330.<br />

Lieckfeldt, E., Kullnig, C., Samuels, G. J. & Kubicek, C. P.<br />

(2000). Sexually competent, sucrose- and nitrateassimilating<br />

strains of Hypocrea jecorina (Trichoderma<br />

reesei) from South American soils. Mycologia, 92,<br />

374 380.<br />

Liew, E. C. Y., Aptroot, A. & Hyde, K. D. (2000).<br />

Phylogenetic significance of the pseudoparaphyses<br />

in Loculoascomycete taxonomy. Molecular<br />

Phylogenetics and Evolution, 16, 392 402.<br />

Lilley, J. H. & Roberts, R. J. (1997). Pathogenicity and<br />

culture studies comparing the Aphanomyces involved<br />

in epizootic ulcerative syndrome (EUS) with<br />

other similar fungi. Journal of Fish Diseases, 20,<br />

135 144.<br />

Limpert, E., Godet, F. & Müller, K. (1999). Dispersal of<br />

cereal mildews across Europe. Agricultural and Forest<br />

Meteorology, 97, 293 308.<br />

Lincoff, G. & Mitchel, D. J. (1977). Toxic and Hallucinogenic<br />

Mushroom Poisoning. New York: Van Nostrand<br />

Reinhold.<br />

Linder, D. H. (1929). A monograph of the helicosporous<br />

fungi imperfecti. Annals of the Missouri Botanical<br />

Gardens, 16, 227 388.<br />

Linderman, R. G. (1997). Vesicular arbuscular mycorrhizal<br />

VAM) fungi. In The Mycota VB: Plant<br />

Relationships, ed. G. C. Carroll & P. Tudzynski. Berlin:<br />

Springer-Verlag, pp. 117 28.<br />

Line, R. F. (2002). Stripe rust of wheat and barley in<br />

North America: a retrospective his<strong>to</strong>rical review.<br />

Annual Review of Phy<strong>to</strong>pathology, 40, 75 118.<br />

Line, R. F. & Griffith, C. S. (2001). Research on the<br />

epidemiology of stem rust of wheat during the Cold<br />

War. In Stem Rust of Wheat. From Ancient Enemy <strong>to</strong><br />

Modern Foe, ed. P. D. Peterson. St. Paul, USA: APS Press,<br />

pp. 83 118.<br />

Lines, C. E. M., Ratcliffe, R. G., Rees, T. A. V. &<br />

Southon, T. E. (1989). A 13 C NMR study of<br />

pho<strong>to</strong>synthate transport and metabolism in the<br />

lichen Xanthoria calcicola Oxner. New Phy<strong>to</strong>logist,<br />

111, 447 456.<br />

Lingappa, B. T. (1958a). Development and cy<strong>to</strong>logy of<br />

the evanescent prosori of Synchytrium brownii Karling.<br />

American Journal of Botany, 45, 116 123.


REFERENCES<br />

763<br />

Lingappa, B. T. (1958b). The cy<strong>to</strong>logy of development<br />

and germination of resting spores of Synchytrium<br />

brownii. American Journal of Botany, 45, 613 620.<br />

Lingle, W. L., Clay, R. P. & Porter, D. (1992).<br />

Ultrastructural analysis of basidiosporogenesis in<br />

Panellus stypticus. Canadian Journal of Botany, 70,<br />

2017 2027.<br />

Lippincott, J. & Li, R. (1998). Sequential assembly of<br />

myosin II, an IQGAP-like protein, and filamen<strong>to</strong>us<br />

actin <strong>to</strong> a ring structure involved in budding yeast<br />

cy<strong>to</strong>kinesis. Journal of Cell Biology, 140, 355 366.<br />

List, P. H. & Freund, B. (1968). Geruchss<strong>to</strong>ffe der<br />

Stinkmorchel, Phallus impudicus L. Planta Medica<br />

(Supplement), 123 132.<br />

Littlefield, L. J. & Heath, M. C. (1979). Ultrastructure of<br />

Rust <strong>Fungi</strong>. New York: Academic Press.<br />

Liu, Y. (2003). Molecular mechanisms of entrainment in<br />

the Neurospora circadian clock. Journal of Biological<br />

Rhythms, 18, 195 205.<br />

Liu, Y. J. & Hall, B. D. (2004). Body plan evolution of<br />

ascomycetes, as inferred from an RNA polymerase II<br />

phylogeny. Proceedings of the National Academy of<br />

Sciences of the USA, 101, 4507 4512.<br />

Liu, Y. J., Whelen, S. & Hall, B. D. (1999). Phylogenetic<br />

relationships among ascomycetes: evidence from an<br />

RNA polymerase II subunit. Molecular Biology and<br />

Evolution, 16, 1799 1808.<br />

Liu, Z.-Y., Liang, Z.-Q., Whalley, A. J. S., Yao, Y.-J. & Liu, A.-<br />

Y. (2001). Cordyceps brittlebankisoides, a new pathogen<br />

of grubs and its anamorph Metarhizium anisopliae var<br />

majus. Journal of Invertebrate Pathology, 78, 178 182.<br />

Liu, Z.-Y., Liang, Z.-Q., Liu, A.-Y., Yao, Y.-J., Hyde, K. D. & Yu,<br />

Z.-N. (2002). Molecular evidence for teleomorphanamorph<br />

connections in Cordyceps based on ITS-5.8S<br />

rDNA sequences. Mycological Research, 106, 1100 1108.<br />

Lizoň, P., Iturriaga, T. & Korf, R. P. (1998). A preliminary<br />

Discomycete flora of Macaronesia: Part 18, Leotiales.<br />

Mycotaxon, 67, 73 83.<br />

LoBuglio, K. F., Pitt, J. I. & Taylor, J. W. (1993).<br />

Phylogenetic analysis of two ribosomal DNA regions<br />

indicates multiple independent losses of a sexual<br />

Talaromyces state among asexual Penicillium species in<br />

subgenus Biverticillium. Mycologia, 85, 592 604.<br />

Lockhart, S. R., Daniels, K. J., Zhao, R., Wessels, D. & Soll,<br />

D. R. (2003). Cell biology of mating in Candida<br />

albicans. Eukaryotic Cell, 2, 49 61.<br />

Loeffler, R. S. T., Butters, J. A. & Hollomon, D. W. (1992).<br />

The sterol composition of powdery mildews.<br />

Phy<strong>to</strong>chemistry, 31, 1561 1563.<br />

Lohmeyer, M. & Tudzynski, P. (1997). Claviceps alkaloids.<br />

In Fungal Biotechnology, ed. T. Anke. Weinheim,<br />

Germany: Chapman & Hall, pp. 173 85.<br />

Long, P. E. & Jacobs, L. (1974). Aseptic fruiting of the<br />

cultivated mushroom Agaricus bisporus. Transactions of<br />

the British Mycological Society, 63, 99 107.<br />

Long, R. M., Singer, R. H., Meng, X., Gonzalez, I.,<br />

Nasmyth, K. & Jansen, R. P. (1997). Mating type<br />

switching in yeast controlled by asymmetric localization<br />

of ASH1 mRNA. Science, 277, 383 387.<br />

López-Franco, R., Howard, R. J. & Bracker, C. E. (1995).<br />

Satellite Spitzenkörper in growing hyphal tips.<br />

Pro<strong>to</strong>plasma, 188, 85 103.<br />

Lorin, S., Dufour, E., Boulay, J., Begel, O., Marsy, S. &<br />

Sainsard-Chanet, A. (2001). Overexpression of the<br />

alternative oxidase res<strong>to</strong>res senescence and fertility<br />

in a long-lived respiration-deficient mutant of Podospora<br />

anserina. Molecular Microbiology, 42, 1259 1267.<br />

Lösel, D. M. (1988). Fungal lipids. In Microbial Lipids 1,<br />

ed. C. Ratledge & S. G. Wilkinson. London: Academic<br />

Press, pp. 699 806.<br />

Lott, T. J., Holloway, B. P., Logan, D. A., Fundyga, R. &<br />

Arnold, J. (1999). Towards understanding the evolution<br />

of the human commensal yeast Candida albicans.<br />

Microbiology, 145, 1137 1143.<br />

Louis, C., Girard, M., Kuhl, G. & Lopez-Ferber, M. (1996).<br />

Persistence of Botrytis cinerea in its vec<strong>to</strong>r Drosophila<br />

melanogaster. Phy<strong>to</strong>pathology, 86, 934 939.<br />

Lovett, J. S. (1975). Growth and differentiation of the<br />

water mold Blas<strong>to</strong>cladiella emersonii. Bacteriological<br />

Reviews, 39, 345 404.<br />

Lowe, S. E., Griffith, G. G., Milne, A., Theodorou, M. K. &<br />

Trinci, A. P. J. (1987a). Life cycle and growth kinetics<br />

of an anaerobic rumen fungus. Journal of General<br />

Microbiology, 133, 1815 1827.<br />

Lowe, S. E., Theodorou, M. K. & Trinci, A. P. J. (1987b).<br />

Isolation of anaerobic fungi from saliva and faeces of<br />

sheep. Journal of General Microbiology, 133, 1829 1834.<br />

Lowy, B. (1952). The genus Auricularia. Mycologia, 44,<br />

656 692.<br />

Lowry, D. S., Fisher, K. E. & Roberson, R. W. (1998).<br />

Establishment and maintenance of nuclear position<br />

during zoospore formation in Allomyces macrogynus.<br />

Fungal Genetics and Biology, 24, 34 44.<br />

Lowry, D. S., Fisher, K. E. & Roberson, R. W. (2004).<br />

Functional necessity of the cy<strong>to</strong>skele<strong>to</strong>n during<br />

cleavage membrane development and zoosporogenesis<br />

in Allomyces macrogynus. Mycologia, 96, 211 218.<br />

Lowry, R. J. & Sussman, A. S. (1958). Wall structure of<br />

ascospores of Neurospora tetrasperma. American Journal<br />

of Botany, 45, 397 403.<br />

Lowry, R. J. & Sussman, A. S. (1968). Ultrastructural<br />

changes during germination of ascospores of<br />

Neurospora tetrasperma. Journal of General Microbiology,<br />

51, 403 409.


764 REFERENCES<br />

Lu, B. C. (1964). Polyploidy in the basidiomycete Cyathus<br />

stercoreus. American Journal of Botany, 51, 343 347.<br />

Lu, G. (2003). Engineering Sclerotinia sclerotiorum resistance<br />

in oilseed crops. African Journal of Biotechnology,<br />

2, 509 516.<br />

Lü, H. & McLaughlin, D. J. (1991). Ultrastructure of the<br />

septal pore apparatus and early septum initiation in<br />

Auricularia auricula-judae. Mycologia, 83, 322 334.<br />

Lu, S.-H. (1973). Effect of calcium on fruiting of Cyathus<br />

stercoreus. Mycologia, 65, 329 334.<br />

Lucarotti, C. J. (1981). Zoospore ultrastructure in<br />

Nowakowskiella elegans and Cladochytrium replicatum<br />

(Chytridiales). Canadian Journal of Botany, 59,<br />

137 148.<br />

Lucarotti, C. J. (1987). Cladochytrium replicatum. In<br />

Zoosporic <strong>Fungi</strong> in Teaching and Research, ed. M. S. Fuller<br />

& A. Jaworski. Athens, USA: Southeastern Publishing<br />

Corporation, pp. 20 1.<br />

Lucarotti, C. J. & Wilson, C. M. (1987). Nowakowskiella<br />

elegans. In Zoosporic <strong>Fungi</strong> in Teaching and Research, ed.<br />

M. S. Fuller & A. Jaworski. Athens, USA: Southeastern<br />

Publishing Corporation, pp. 22 3.<br />

Lucas, J. A., Shat<strong>to</strong>ck, R. C., Shaw, D. S. & Cooke, L. R.,<br />

eds. (1991). Phy<strong>to</strong>phthora. Cambridge: Cambridge<br />

University Press.<br />

Lucas, J. A., Dyer, P. S. & Murray, T. D. (2000).<br />

Pathogenicity, host-specificity, and population biology<br />

of Tapesia spp., causal agents of eyespot<br />

disease of cereals. Advances in Botanical Research, 33,<br />

225 258.<br />

Ludwig-Müller, J. (1999). Plasmodiophora brassicae, the<br />

causal agent of clubroot disease: a review on<br />

molecular and biochemical events in pathogenesis.<br />

Zeitschrift für Pflanzenkrankheiten und Pflanzenschutz,<br />

106, 109 127.<br />

Ludwig-Müller, J., Bendel, U., Thermann, P., et al. (1993).<br />

Concentrations of indole-3-acetic acid in plants of<br />

<strong>to</strong>lerant and susceptible varieties of Chinese cabbage<br />

infected with Plasmodiophora brassicae Woron. New<br />

Phy<strong>to</strong>logist, 125, 763 769.<br />

Ludwig-Müller, J., Epstein, E. & Hilgenberg, W. (1996).<br />

Auxin-conjugate hydrolysis in Chinese cabbage:<br />

characterization of an amidohydrolase and its role<br />

during the clubroot disease. Physiologia Plantarum, 97,<br />

627 634.<br />

Ludwig-Müller, J., Bennett, R. N., Kiddle, G., et al. (1999).<br />

The host range of Plasmodiophora brassicae and its<br />

relationship <strong>to</strong> endogenous glucosinolate content.<br />

New Phy<strong>to</strong>logist, 141, 443 458.<br />

Lugones, L. G., Boscher, J. S., Scholtmeyer, K., De Vries,<br />

O. M. H. & Wessels, J. G. H. (1996). An abundant<br />

hydrophobin (ABH1) forms hydrophobic rodlet layers<br />

in Agaricus bisporus fruiting bodies. Microbiology, 142,<br />

1321 1329.<br />

Lugones, L. G., Wösten, H. A. B., Birkenkamp, K. U., et al.<br />

(1999). Hydrophobins line air channels in fruiting<br />

bodies of Schizophyllum commune and Agaricus bisporus.<br />

Mycological Research, 103, 635 640.<br />

Lumbsch, H. T. & Lindemuth, R. (2001). Major lineages<br />

of Dothideomycetes (Ascomycota) inferred from SSU and<br />

LSU rDNA sequences. Mycological Research, 105,<br />

901 908.<br />

Lumbsch, H. T., Schmitt, I., Döring, H. & Wedin, M.<br />

(2001). Molecular systematics supports the recognition<br />

of an additional order of Ascomycota: the<br />

Agyriales. Mycological Research, 105, 16 23.<br />

Lumbsch, H. T., Schmitt, I., Palice, Z., et al. (2004).<br />

Supraordinal phylogenetic relationships of<br />

Lecanoromycetes based on a Bayesian analysis of<br />

combined nuclear and mi<strong>to</strong>chondrial sequences.<br />

Molecular Phylogenetics and Evolution, 31, 822 832.<br />

Lumsden, R. D. (1992). Mycoparasitism of soilborne<br />

plant pathogens. In The Fungal Community. Its<br />

Organization and Role in the Ecosystem, 2nd edn, ed.<br />

G. C. Carroll & D. T. Wicklow. New York: Marcel<br />

Dekker Inc., pp. 275 93.<br />

Lundqvist, N. (1972). Nordic Sordariaceae s. lat. Symbolae<br />

Botanicae Upsalienses, 20, 1 374.<br />

Lundqvist, N., Mahoney, D. P., Bell, A. & Lorenzo, L. E.<br />

(1999). Podospora austrohemisphaerica, a new heterothallic<br />

ascomycete from dung. Mycologia, 91,<br />

405 415.<br />

Lunney, C. Z. & Bland, C. E. (1976). An ultrastructural<br />

study of zoosporogenesis in Pythium proliferum de<br />

Bary. Pro<strong>to</strong>plasma, 88, 85 100.<br />

Luttrell, E. S. (1951). Taxonomy of the Pyrenomycetes.<br />

University of Missouri Studies, 24(3), 1 20.<br />

Luttrell, E. S. (1955). The ascostromatic ascomycetes.<br />

Mycologia, 47, 511 532.<br />

Luttrell, E. S. (1963). Taxonomic criteria in<br />

Helminthosporium. Mycologia, 55, 643 674.<br />

Luttrell, E. S. (1965). Paraphysoids, pseudoparaphyses<br />

and apical paraphyses. Transactions of the British<br />

Mycological Society, 48, 135 144.<br />

Luttrell, E. S. (1973). Loculoascomycetes. In The <strong>Fungi</strong>: An<br />

Advanced Treatise IVA, ed. G. C. Ainsworth, F. K.<br />

Sparrow & A. S. Sussman. New York: Academic Press,<br />

pp. 135 219.<br />

Luttrell, E. S. (1977). The disease cycle and fungus host<br />

relationships in dallisgrass ergot. Phy<strong>to</strong>pathology, 67,<br />

1461 1468.<br />

Luttrell, E. S. (1980). Host parasite relationships and<br />

development of the ergot sclerotium in Claviceps<br />

purpurea. Canadian Journal of Botany, 58, 942 958.


REFERENCES<br />

765<br />

Luttrell, E. S. (1981). The pyrenomycete centrum. In<br />

Ascomycete Systematics. The Luttrellian Concept, ed.<br />

D. R. Reynolds. New York: Springer-Verlag,<br />

pp. 124 37.<br />

Lutzoni, F., Pagel, M. & Reeb, B. (2001). Major fungal<br />

lineages are derived from lichen symbiotic ances<strong>to</strong>rs.<br />

Nature, 411, 937 940.<br />

Ma, L.-J., Rogers, S. O., Catranis, C. M. & Starmer, W. T.<br />

(2000). Detection and characterization of ancient<br />

fungi entrapped in glacial ice. Mycologia, 92,<br />

286 295.<br />

Macfarlane, I. (1952). Fac<strong>to</strong>rs affecting the survival of<br />

Plasmodiophora brassicae Wor. in the soil and its<br />

assessment by a host test. Annals of Applied Biology, 39,<br />

239 256.<br />

Macfarlane, I. (1968). Problems in the systematics of the<br />

Olpidiaceae. In Marine Mykologie (Symposium über<br />

Niedere Pilze im Küstenbereich), ed. A. Gaertner.<br />

Bremerhaven: Institut für Meeresforschung,<br />

pp. 39 58.<br />

Macfarlane, I. (1970). Germination of resting spores of<br />

Plasmodiophora brassicae. Transactions of the British<br />

Mycological Society, 55, 97 112.<br />

Macfarlane, T. D., Kuo, J. & Hil<strong>to</strong>n, R. N. (1978).<br />

Structure of the giant sclerotium of Polyporus<br />

mylittae. Transactions of the British Mycological Society,<br />

71, 359 365.<br />

MacHardy, W. E., Gadoury, D. M. & Gessler, C. (2001).<br />

Parasitic and biological fitness of Venturia inaequalis:<br />

relationship <strong>to</strong> disease management strategies. Plant<br />

Disease, 85, 1036 1051.<br />

Machlis, L. (1972). The coming of age of sex hormones<br />

in plants. Mycologia, 64, 235 247.<br />

Macko, V., Staples, R. C., Gershon, H. & Renwick, J. A. A.<br />

(1970). Self-inhibi<strong>to</strong>r of bean rust uredospores:<br />

methyl 3,4-dimethoxycinnamate. Science, 170,<br />

539 540.<br />

Maclean, D. J. (1982). Axenic culture and metabolism of<br />

rust fungi. In The Rust <strong>Fungi</strong>, ed. K. J. Scott & A. K.<br />

Chakravorty. London: Academic Press, pp. 37 84.<br />

MacLeod, D. M., Müller-Kögler, E. & Wilding, N. (1976).<br />

En<strong>to</strong>mophthora species with E. muscae-like conidia.<br />

Mycologia, 68, 1 29.<br />

Madelin, M. F. (1984). Presidential address. Myxomycete<br />

data of ecological significance. Transactions of the<br />

British Mycological Society, 83, 1 19.<br />

Magee, B. B. & Magee, P. T. (2000). Induction of mating<br />

in Candida albicans by construction of MTLa and<br />

MTLa strains. Science, 289, 310 313.<br />

Magliani, W., Conti, S., Gerloni, M., Ber<strong>to</strong>lotti, D. &<br />

Polonelli, L. (1997). Yeast killer systems. Clinical<br />

Microbiology Reviews, 10, 369 400.<br />

Magliani, W., Conti, S., Arseni, S., et al. (2005). Antibodymediated<br />

protective immunity in fungal infections.<br />

New Microbiologica, 28, 299 309.<br />

Maheshwari, R. (1999). Microconidia of Neurospora<br />

crassa. Fungal Genetics and Biology, 26, 1 18.<br />

Maheshwari, R., Bharadwaj, G. & Bhat, M. K. (2000).<br />

Thermophilic fungi: their physiology and enzymes.<br />

Microbiology and Molecular Biology Reviews, 64,<br />

461 488.<br />

Mahfoud, R., Maresca, M., Garmy, N. & Fantini, J. (2002).<br />

The myco<strong>to</strong>xin patulin alters the barrier function of<br />

the intestinal epithelium: mechanism of action of<br />

the <strong>to</strong>xin and protective effects of glutathione.<br />

Toxicology and Applied Pharmacology, 181, 209 218.<br />

Mahr, F. A. & Bruzzese, E. (1998). The effect of<br />

Phragmidium violaceum (Shultz) Winter (Uredinales)<br />

on Rubus fruticosus L. agg. in south-eastern Vic<strong>to</strong>ria.<br />

Plant Protection Quarterly, 13, 182 185.<br />

Mai, S. H. (1976). Morphological studies in Podospora<br />

anserina. American Journal of Botany, 63, 821 825.<br />

Mai, S. H. (1977). Morphological studies in Sordaria<br />

fimicola and Gelasinospora longispora. American Journal<br />

of Botany, 64, 489 495.<br />

Maia, J. C. D. (1994). Hexosamine and cell wall biogenesis<br />

in the aquatic fungus Blas<strong>to</strong>cladiella emersonii.<br />

FASEB Journal, 8, 848 853.<br />

Maia, L. C., Yano, A. M. & Kimbrough, J. W. (1996).<br />

Species of Ascomycota forming ec<strong>to</strong>mycorrhizae.<br />

Mycotaxon, 57, 371 390.<br />

Maier, W., Begerow, D., Weiss, M. & Oberwinkler, F.<br />

(2003). Phylogeny of the rust fungi: an approach<br />

using nuclear large subunit ribosomal DNA<br />

sequences. Canadian Journal of Botany, 81, 12 23.<br />

Mains, E. B. (1957). Species of Cordyceps parasitic on<br />

Elaphomyces. Bulletin of the Torrey Botanical Club, 84,<br />

243 251.<br />

Mainwaring, H. R. (1972). The fine structure of ascospore<br />

wall formation in Sordaria fimicola. Archiv für<br />

Mikrobiologie, 81, 126 135.<br />

Maitland, D. P. (1994). A parasitic fungus infecting<br />

yellow dung flies manipulates host perching behaviour.<br />

Proceedings of the Royal Society of London, B 258,<br />

187 193.<br />

Malcolmson, J. F. (1969). Races of Phy<strong>to</strong>phthora infestans<br />

occurring in Great Britain. Transactions of the British<br />

Mycological Society, 53, 417 423.<br />

Malcolmson, J. F. (1970). Vegetative hybridity in<br />

Phy<strong>to</strong>phthora infestans. Nature, 225, 971 972.<br />

Maleck, K., Neuenschwander, U., Cade, R. M., et al.<br />

(2002). Isolation and characterization of broadspectrum<br />

disease-resistant Arabidopsis mutants.<br />

Genetics, 160, 1661 1671.


766 REFERENCES<br />

Malik, M. M. S. (1974). Nuclear behaviour during teliospore<br />

germination in Ustilago tritici and U. nuda.<br />

Pakistan Journal of Botany, 6, 59 63.<br />

Mallett, K. J. & Harrison, I. M. (1988). The mating system<br />

of the fairy ring fungus Marasmius oreades and the<br />

genetic relationship of fairy rings. Canadian Journal of<br />

Botany, 66, 1111 1116.<br />

Malloch, D. W. (1987). The evolution of mycorrhizae.<br />

Canadian Journal of Plant Pathology, 9, 398 402.<br />

Malloch, D. & Blackwell, M. (1993). Dispersal biology of<br />

ophios<strong>to</strong>ma<strong>to</strong>id fungi. In Cera<strong>to</strong>cystis and Ophios<strong>to</strong>ma:<br />

Taxonomy, Ecology and Pathology, ed. M. J. Wingfield,<br />

K. A. Seifert & J. F. Webber. St. Paul, USA: APS Press,<br />

pp. 195 206.<br />

Maloy, O. C. (1997). White pine blister rust control in<br />

North America: a case his<strong>to</strong>ry. Annual Review of<br />

Phy<strong>to</strong>pathology, 35, 87 109.<br />

Mamoun, M. & Olivier, J.-M. (1996). Receptivity of<br />

cloned hazels <strong>to</strong> artificial ec<strong>to</strong>mycorrhizal infection<br />

by Tuber melanosporum and symbiotic competi<strong>to</strong>rs.<br />

Mycorrhiza, 6, 15 19.<br />

Manion, P. & Griffin, D. H. (1986). Sixty-five years of<br />

research on Hypoxylon canker of aspen. Plant Disease,<br />

70, 803 808.<br />

Manners, D. J. & Meyer, M. T. (1977). The molecular<br />

structures of some glucans from the cell walls of<br />

Schizosaccharomyces pombe. Carbohydrate Research, 57,<br />

189 203.<br />

Manners, J. M. (1989). The host haus<strong>to</strong>rium interface<br />

in powdery mildews. Australian Journal of Plant<br />

Physiology, 16, 45 52.<br />

Manocha, M. S. (1965). Fine structure of the Agaricus<br />

carpophore. Canadian Journal of Botany, 43,<br />

1329 1333.<br />

Manocha, M. S. (1975). Host parasite relations in a<br />

mycoparasite. III. Morphological and biochemical<br />

differences in the parasitic and axenic-culture<br />

spores of Pip<strong>to</strong>cephalis virginiana. Mycologia, 67,<br />

383 391.<br />

Manocha, M. S. (1981). Host specificity and mechanisms<br />

of resistance in a mycoparasitic system. Physiological<br />

Plant Pathology, 18, 257 265.<br />

Manocha, M. S. & Deven, J. M. (1975). Host parasite<br />

relations in a mycoparasite. IV. A correlation<br />

between the levels of gamma-linolenic acid and<br />

parasitism in Pip<strong>to</strong>cephalis virginiana. Mycologia, 67,<br />

1148 1157.<br />

Manocha, M. S. & Golesorkhi, R. (1981). Host parasite<br />

relations in a mycoparasite. 7. Light and scanning<br />

electron microscopy of interactions of Pip<strong>to</strong>cephalis<br />

virginiana with host and non-host species. Mycologia,<br />

73, 976 987.<br />

Manocha, M. S. & Lee, K. Y. (1971). Host parasite<br />

relations in a mycoparasite. I. Fine structure of host,<br />

parasite and their interface. Canadian Journal of<br />

Botany, 49, 1677 1681.<br />

Manocha, M. S. & McCullough, C. M. (1985). Suppression<br />

of host cell wall synthesis at penetration sites in a<br />

compatible interaction with a mycoparasite.<br />

Canadian Journal of Botany, 63, 967 973.<br />

Manocha, M. S. & Zhonghua, Z. (1997).<br />

Immunocy<strong>to</strong>chemical and cy<strong>to</strong>chemical localization<br />

of chitinase and chitin in infected hosts of a<br />

biotrophic mycoparasite Pip<strong>to</strong>cephalis virginiana.<br />

Mycologia, 89, 185 194.<br />

Manocha, M. S., Chen, Y. & Rao, N. (1990). Involvement<br />

of cell surface sugars in recognition, attachment and<br />

appressorium formation by a mycoparasite.<br />

Canadian Journal of Microbiology, 36, 771 778.<br />

Manocha, M. S., Xiong, D. & Govindasamy, V. (1997).<br />

Isolation and partial characterization of a complementary<br />

protein from the mycoparasite Pip<strong>to</strong>cephalis<br />

virginiana that specifically binds <strong>to</strong> two glycoproteins<br />

at the host cell surface. Canadian Journal of<br />

Microbiology, 43, 625 632.<br />

Mantegani, S., Brambilla, E. & Varasi, M. (1999).<br />

Ergoline derivatives: recep<strong>to</strong>r affinity and selectivity.<br />

Farmaco, 54, 288 296.<br />

Mantiri, F. R., Samuels, G. J., Rahe, J. E. & Honda, B. M.<br />

(2001). Phylogenetic relationships of Neonectria<br />

species having Cylindrocarpon anamorphs inferred<br />

from mi<strong>to</strong>chondrial ribosomal DNA sequences.<br />

Canadian Journal of Botany, 79, 334 340.<br />

Maras, M., Saelens, X., Laroy, W., et al. (1997). In vitro<br />

conversion of the carbohydrate moiety of fungal<br />

glycoproteins <strong>to</strong> mammalian-type oligosaccharides.<br />

Evidence for N-acetylglucosaminyltransferase-Iaccepting<br />

glycans from Trichoderma reesei. European<br />

Journal of Biochemistry, 249, 701 707.<br />

Marchant, R. & Wessels, J. G. H. (1974). An ultrastructural<br />

study of septal dissolution in Schizophyllum<br />

commune. Archiv für Mikrobiologie, 96, 175 182.<br />

Marek, L. E. (1984). Light affects in vitro development of<br />

gametangia and sporangia of Monoblepharis macrandra<br />

(Chytridiomycetes, Monoblepharidales).<br />

Mycologia, 76, 420 425.<br />

Margulis, L., Corliss, J. O., Melkonian, M. & Chapman,<br />

D. J., eds. (1990). Handbook of the Pro<strong>to</strong>ctista. Bos<strong>to</strong>n:<br />

Jones & Bartlett.<br />

Markham, P. (1994). Occlusions of the septal pores of<br />

filamen<strong>to</strong>us fungi. Mycological Research, 98,<br />

1089 1106.<br />

Markham, P. & Collinge, A. J. (1987). Woronin bodies of<br />

filamen<strong>to</strong>us fungi. FEMS Microbiology Reviews, 46,1 11.


REFERENCES<br />

767<br />

Markovich, N. A. & Kononova, G. L. (2003). Lytic<br />

enzymes of Trichoderma and their role in plant<br />

defense from fungal diseases: a review. Applied<br />

Biochemistry and Microbiology, 39, 341 351.<br />

Marquina, D., San<strong>to</strong>s, A. & Peinado, J. M. (2002). Biology<br />

of killer yeasts. International Microbiology, 5, 65 71.<br />

Marshall, G. M. (1960). The incidence of certain seedborne<br />

diseases in commercial seed-samples. IV. Bunt<br />

of wheat, Tilletia caries (DC.) Tul. V. Earcockles of<br />

wheat, Anguina tritici (Stein.) Filipjev. Annals of Applied<br />

Biology, 48, 34 38.<br />

Martin, E. (1940). The morphology and cy<strong>to</strong>logy of<br />

Taphrina deformans. American Journal of Botany, 27,<br />

743 751.<br />

Martin, F., Díez, J., Dell, B. & Delaruelle, C. (2002).<br />

Phylogeography of the ec<strong>to</strong>mycorrhizal Pisolithus<br />

species as inferred from nuclear ribosomal DNA ITS<br />

sequences. New Phy<strong>to</strong>logist, 153, 345 357.<br />

Martin, G. W. (1945). The classification of the<br />

Tremellales. Mycologia, 37, 527 542.<br />

Martin, H. D., Ruck, C., Schmidt, M., et al. (1999).<br />

Chemistry of carotenoid oxidation and free radical<br />

reactions. Pure and Applied Chemistry, 71, 2253 2262.<br />

Martin, J. F., Gutiérrez, S. & Demain, A. L. (1997).<br />

Antibiotics: b-lactams. In Fungal Biotechnology, ed. T.<br />

Anke. Weinheim, Germany: Chapman & Hall,<br />

pp. 91 127.<br />

Martin, M., Gay, J. L. & Jackson, G. V. H. (1976). Electron<br />

microscopic study of developing and mature cleis<strong>to</strong>thecia<br />

of Sphaerotheca mors-uvae. Transactions of the<br />

British Mycological Society, 66, 473 487.<br />

Martín, M. P. (1996). The Genus Rhizopogon in Europe.<br />

Barcelona: Societat Catalana de Micologia.<br />

Martinez, M., López-Ribot, J., Kirkpatrick, W. R., et al.<br />

(2002). Replacement of Candida albicans with C.<br />

dubliniensis in human immunodeficiency virus-treated<br />

patients with oropharyngeal candidiasis treated<br />

with fluconazole. Journal of Clinical Microbiology, 40,<br />

3135 3139.<br />

Martínez-Espinoza, A. D., García-Pedrajas, M. D. & Gold,<br />

S. E. (2002). The Ustilaginales as plant pests and<br />

model systems. Fungal Genetics and Biology, 35, 1 20.<br />

Marx, D. H. (1977). Tree host range and world distribution<br />

of the ec<strong>to</strong>mycorrhizal fungus Pisolithus tinc<strong>to</strong>rius.<br />

Canadian Journal of Microbiology, 23, 217 223.<br />

Massicotte, H. B., Peterson, R. L. & Ashford, A. E. (1987).<br />

On<strong>to</strong>geny of Eucalyptus pilularis Pisolithus tinc<strong>to</strong>rius<br />

ec<strong>to</strong>mycorrhiza. I. Light microscopy and scanning<br />

electron microscopy. Canadian Journal of Botany, 65,<br />

1927 1939.<br />

Masterman, R., Ross, R., Mesce, K. & Spivak, M. (2001).<br />

Olfac<strong>to</strong>ry and behavioural response thresholds <strong>to</strong><br />

odors of diseased brood differ between hygienic and<br />

non-hygienic honey bees (Apis mellifera L.). Journal of<br />

Comparative Physiology A, 187, 441 452.<br />

Masuch, G. (1993). Biologie der Flechten. Heidelberg and<br />

Wiesbaden: Quelle & Meyer.<br />

Masuhara, G., Katsuya, K. & Yamaguchi, K. (1993).<br />

Potential for symbiosis of Rhizoc<strong>to</strong>nia solani and<br />

binucleate Rhizoc<strong>to</strong>nia with seeds of Spiranthes sinensis<br />

var. amoena in vitro. Mycological Research, 97, 746 752.<br />

Mata, J. & Nurse, P. (1998). Discovering the poles in<br />

yeast. Trends in Cell Biology, 8, 163 167.<br />

Mather, K. & Jinks, J. L. (1958). Cy<strong>to</strong>plasm in sexual<br />

reproduction. Nature, 182, 1188 1190.<br />

Mathew, K. T. (1961). Morphogenesis of mycelial<br />

strands in the cultivated mushroom, Agaricus<br />

bisporus. Transactions of the British Mycological Society,<br />

44, 285 290.<br />

Mathre, D. E. (1996). Dwarf bunt: politics, identification,<br />

and biology. Annual Review of Phy<strong>to</strong>pathology, 34,<br />

67 85.<br />

Ma<strong>to</strong>ssian, M. K. (1989). Poisons of the Past: Molds,<br />

Epidemics and His<strong>to</strong>ry. New Haven: Yale University<br />

Press.<br />

Matsumo<strong>to</strong>, C., Kageyama, K., Suga, H. & Hyakumachi,<br />

M. (1999). Phylogenetic relationships of Pythium<br />

species based on ITS and 5.8S sequences. Mycoscience,<br />

40, 321 331.<br />

Mau, J. L., Beelman, R. B. & Ziegler, G. R. (1992). Effect of<br />

10-oxo-trans-B decenoic acid on the growth of<br />

Agaricus bisporus. Phy<strong>to</strong>chemistry, 31, 4059 4064.<br />

Mayer, A. M., Staples, R. C. & Gil-ad, N. L. (2001).<br />

Mechanisms of survival of necrotrophic fungal plant<br />

pathogens in hosts expressing the hypersensitive<br />

response. Phy<strong>to</strong>chemistry, 58, 33 41.<br />

Mazars, C., Canivenc, E., Rossignol, M. & Auriol, P.<br />

(1991). Production of phomozin in sunflower<br />

following artificial inoculation with Phomopsis<br />

helianthi. Plant Science, 75, 155 160.<br />

Mazzola, M. (2002). Mechanisms of natural soil<br />

suppressiveness <strong>to</strong> soilborne diseases. An<strong>to</strong>nie van<br />

Leeuwenhoek, 81, 557 564.<br />

McAfee, B. J. & Fortin, J. A. (1986). Competitive interactions<br />

of ec<strong>to</strong>mycorrhizal mycobionts under field<br />

conditions. Canadian Journal of Botany, 64, 848 852.<br />

McCabe, D. E., Humber, R. A. & Soper, R. S. (1984).<br />

Observation and interpretation of nuclear reductions<br />

during maturation and germination of en<strong>to</strong>mophthoralean<br />

resting spores. Mycologia, 76,<br />

1104 1107.<br />

McCartney, H. A., Schmechtel, D. & Lacey, M. E. (1993).<br />

Aerodynamic diameter of conidia of Alternaria<br />

species. Plant Pathology, 42, 280 286.


768 REFERENCES<br />

McCreadie, J. W., Beard, C. E. & Adler, P. H. (2005).<br />

Context-dependent symbiosis between black flies<br />

(Diptera: Simuliidae) and trichomycete fungi<br />

(Harpellales: Legeriomycetaceae). Oikos, 108,<br />

362 370.<br />

McDaniell, L. L. & Hindal, D. F. (1982). Spore swelling,<br />

germination, and germ tube formation in axenic<br />

culture among four species of Pip<strong>to</strong>cephalis. Mycologia,<br />

74, 271 274.<br />

McDowell, J. M., Cuzick, A., Can, C., et al. (2000). Downy<br />

mildew (Peronospora parasitica) resistance genes in<br />

Arabidopsis vary in functional requirements for NDR1,<br />

EDS1, NPR1 and salicylic acid accumulation. Plant<br />

Journal, 22, 523 529.<br />

McGinnis, M. R. (1980). Recent taxonomic developments<br />

and changes in medical mycology. Annual<br />

Review of Microbiology, 34, 109 135.<br />

McGovern, P. E. (2003). Ancient Wine. The Search for the<br />

Origins of Viticulture. Prince<strong>to</strong>n and Oxford: Prince<strong>to</strong>n<br />

University Press.<br />

McKeen, W. E., Mitchell, N. & Smith, R. (1967). The<br />

Erysiphe cichoracearum conidium. Canadian Journal of<br />

Botany, 45, 1489 1496.<br />

McKemy, J. M. & Morgan-Jones, G. (1991). Studies in the<br />

genus Cladosporium sensu la<strong>to</strong>. V. Concerning the type<br />

species, Cladosporium herbarum. Mycotaxon, 42,<br />

307 317.<br />

McKendrick, S. L., Leake, J. R. & Read, D. J. (2000).<br />

Symbiotic germination and development of mycoheterotrophic<br />

plants in nature: transfer of carbon<br />

from ec<strong>to</strong>mycorrhizal Salix repens and Betula pendula<br />

<strong>to</strong> the orchid Corallorhiza trifida through shared<br />

hyphal connections. New Phy<strong>to</strong>logist, 145, 539 548.<br />

McLaughlin, D. J. (1973). Ultrastructure of sterigma<br />

growth and basidiospore formation in Coprinus and<br />

Boletus. Canadian Journal of Botany, 51, 145 150.<br />

McLaughlin, D. J. (1977). Basidiospore initiation and<br />

early development in Coprinus cinereus. American<br />

Journal of Botany, 64, 1 16.<br />

McLaughlin, D. J. (1982). Ultrastructure and cy<strong>to</strong>chemistry<br />

of basidial and basidiospore development. In<br />

Basidium and Basidiocarp: Evolution, Cy<strong>to</strong>logy, Function<br />

and Development, ed. K. Wells & E. K. Wells. New York:<br />

Springer-Verlag, pp. 37 74.<br />

McLaughlin, D. J., Frieders, E. M. & Lü, H. (1995). A<br />

microscopist’s view of heterobasidiomycete phylogeny.<br />

Studies in Mycology, 38, 91 109.<br />

McLaughlin, D. J., McLaughlin, E. G. & Lemke, P. E., eds.<br />

(2001). The Mycota VIIA and VIIB. Berlin: Springer-<br />

Verlag.<br />

McLaughlin, R. J., Wilson, C. L., Chalutz, D., et al. (1990).<br />

Characterization and reclassification of yeasts used<br />

for biological control of postharvest diseases of fruits<br />

and vegetables. Applied and Environmental<br />

Microbiology, 56, 3583 3586.<br />

McMeekin, D. (1960). The role of the oospores of<br />

Peronospora parasitica in downy mildew of crucifers.<br />

Phy<strong>to</strong>pathology, 50, 93 97.<br />

McMillan, M. S., Schwartz, H. F. & Ot<strong>to</strong>, K. L. (2003).<br />

Sexual stage development of Uromyces appendiculatus<br />

and its potential use for disease resistance screening<br />

of Phaseolus vulgaris. Plant Disease, 87, 1133 1138.<br />

McMorris, T. C. (1978). Sex hormones of the aquatic<br />

fungus Achlya. Lipids, 13, 716 722.<br />

McMorris, T. C., Seshadri, R., Weihe, G. R., Arsenault,<br />

G. P. & Barksdale, A. W. (1975). Structure of oogoniol-<br />

1, -2 and -3, steroidal sex hormones of the water<br />

mould Achlya. Journal of the American Chemical Society,<br />

97, 2544 3545.<br />

Meehan, F. & Murphy, H. C. (1946). A new<br />

Helminthosporium blight of oats. Science, 104, 413 414.<br />

Mehta, B. J., Obraztsova, I. N. & Cerdá-Olmedo, E. (2003).<br />

Mutants and intersexual heterokaryons of Blakeslea<br />

trispora for production of b-carotene and lycopene.<br />

Applied and Environmental Microbiology, 69,<br />

4043 4048.<br />

Meier, F. A., Scherrer, S. & Honegger, R. (2002). Faecal<br />

pellets of lichenivorous mites contain viable cells of<br />

the lichen-forming ascomycete Xanthoria parietina<br />

and its green algal pho<strong>to</strong>biont, Trebouxia arboricola.<br />

Biological Journal of the Linnean Society, 76, 259 268.<br />

Meireles, M. C. A., Riet-Correa, F., Fischman, O., et al.<br />

(1993). Cutaneous pythiosis in horses from Brazil.<br />

Mycoses, 36, 139 142.<br />

Meistrich, M., Fork, R. L. & Matricon, J. (1970).<br />

Pho<strong>to</strong>tropism in Phycomyces as investigated by<br />

focused laser radiation. Science, 169, 370 371.<br />

Mendgen, K. (1997). The Uredinales. In The Mycota VB:<br />

Plant Relationships, ed. G. C. Carroll & P. Tudzynski.<br />

Berlin: Springer-Verlag, pp. 79 94.<br />

Mendgen, K. & Deising, H. (1993). Infection structures of<br />

fungal plant pathogens a cy<strong>to</strong>logical and physiological<br />

evaluation. New Phy<strong>to</strong>logist, 124, 193 213.<br />

Mendgen, K. & Hahn, M. (2002). Plant infection and the<br />

establishment of fungal biotrophy. Trends in Plant<br />

Science, 7, 352 356.<br />

Mendoza, L., Villalobos, J., Calleja, C. E. & Solis, A.<br />

(1992). Evaluation of two vaccines for the treatment<br />

of pythiosis insidiosi in horses. Mycopathologia, 119,<br />

89 95.<br />

Mendoza, L., Hernandez, F. & Ajello, L. (1993). Life cycle<br />

of the human and animal oomycete pathogen<br />

Pythium insidiosum. Journal of Clinical Microbiology, 31,<br />

2967 2973.


REFERENCES<br />

769<br />

Menge, J. A. (1984). Inoculum production. In VA<br />

Mycorrhiza, ed. C. L. Powell & D. J. Bagyaraj. Boca<br />

Ra<strong>to</strong>n: CRC Press, pp. 187 203.<br />

Menzies, J. G., Knox, R. E., Nielsen, J. & Thomas, P. L.<br />

(2003). Virulence of Canadian isolates of Ustilago<br />

tritici: 1964 1998, and the use of the geometric rule<br />

in understanding host differential complexity.<br />

Canadian Journal of Plant Pathology, 25, 62 72.<br />

Meredith, D. S. (1965). Violent spore release in<br />

Helminthosporium turcicum. Phy<strong>to</strong>pathology, 55,<br />

1099 1102.<br />

Meredith, D. S. (1966). Diurnal periodicity and violent<br />

liberation of conidia in Epicoccum. Phy<strong>to</strong>pathology, 56,<br />

988.<br />

Merkus, E. (1976). Ultrastructure of the ascospore wall<br />

in Pezizales (Ascomycetes). IV. Morchellaceae,<br />

Helvellaceae, Rhizinaceae, Thelebolaceae and<br />

Sarcoscyphaceae. General discussion. Persoonia, 9,<br />

1 38.<br />

Merz, U. (1997). Microscopical observations of the<br />

primary zoospores of Spongospora subterranea f. sp.<br />

subterranea. Plant Pathology, 46, 670 674.<br />

Metcalf, D. A. & Wilson, C. R. (1999). His<strong>to</strong>logy of<br />

Sclerotium cepivorum infection of onion roots and the<br />

spatial relationships of pectinases in the infection<br />

process. Plant Pathology, 48, 445 452.<br />

Metcalf, D. A. & Wilson, C. R. (2001). The process of<br />

antagonism of Sclerotium cepivorum in white rot<br />

affected onion roots by Trichoderma koningii. Plant<br />

Pathology, 50, 249 257.<br />

Metwalli, A. A. & Shearer, C. A. (1989). Aquatic hyphomycete<br />

communities in clear-cut and wooded areas<br />

of an Illinois stream. Transactions of the Illinois<br />

Academy of Science, 82, 5 16.<br />

Metzenberg, R. L. & Glass, N. L. (1990). Mating type and<br />

mating strategies in Neurospora. Bioessays, 12, 53 59.<br />

Meyer, R. J. & Fuller, M. S. (1985). Structure and<br />

development of hyphal septa in Allomyces. American<br />

Journal of Botany, 72, 1458 1465.<br />

Meyer, S. A., Payne, R. W. & Yarrow, D. (1998). Candida<br />

Berkhout. In The Yeasts: A Taxonomic Study, 4th edn,<br />

ed. C. P. Kurtzman & J. W. Fell. Amsterdam: Elsevier,<br />

pp. 454 573.<br />

Meyer, W. & Gams, W. (2003). Delimitation of<br />

Umbelopsis (Mucorales, Umbelopsidaceae fam. nov.)<br />

based on ITS sequence and RFLP data. Mycological<br />

Research, 107, 339 350.<br />

Miadlikowska, J. & Lutzoni, F. (2004). Phylogenetic<br />

classification of peltigeralean fungi (Peltigerales,<br />

Ascomycota) based on ribosomal RNA small and<br />

large subunits. American Journal of Botany, 91,<br />

449 464.<br />

Micali, C. O. & Smith, M. I. (2003). On the independence<br />

of barrage formation and heterokaryon incompatibility<br />

in Neurospora crassa. Fungal Genetics and Biology,<br />

38, 209 219.<br />

Michaelides, J. & Kendrick, B. (1982). The bubble trap<br />

propagules of Beverwijkella, Helicoon and other aeroaquatic<br />

fungi. Mycotaxon, 14, 247 260.<br />

Michailides, T. J., Guo, L. Y. & Morgan, D. P. (1997).<br />

Fac<strong>to</strong>rs affecting zygosporogenesis in Mucor piriformis<br />

and Gilbertella persicaria. Mycologia, 89, 603 609.<br />

Michelmore, R. W. & Ingram, D. S. (1980).<br />

Heterothallism in Bremia lactucae. Transactions of the<br />

British Mycological Society, 75, 47 56.<br />

Michelot, D. & Melendez-Howell, L. M. (2003). Amanita<br />

muscaria: chemistry, biology, <strong>to</strong>xicology and ethnomycology.<br />

Mycological Research, 107, 131 146.<br />

Michelot, D. & Tebbett, I. (1990). Poisoning by members<br />

of the genus Cortinarius. Mycological Research, 94,<br />

289 298.<br />

Middle<strong>to</strong>n, J. T. (1943). The taxonomy, host range and<br />

geographic distribution of the genus Pythium.<br />

Memoirs of the Torrey Botanical Club, 20, 1 171.<br />

Milburn, J. A. (1970). Cavitation and osmotic potentials<br />

of Sordaria ascospores. New Phy<strong>to</strong>logist, 69, 133 141.<br />

Millay, M. A. & Taylor, T. N. (1978). Chytrid-like fossils of<br />

Pennsylvanian age. Science, 200, 1147 1149.<br />

Miller, C. E. & Dylewski, D. P. (1981). Syngamy and<br />

resting body development in Chytriomyces hyalinus<br />

(Chytridiales). American Journal of Botany, 68, 342 349.<br />

Miller, C. E. & Dylewski, D. P. (1987). Chytriomyces<br />

hyalinus. In Zoosporic <strong>Fungi</strong> in Teaching and Research, ed.<br />

M. S. Fuller & A. Jaworski. Athens, USA: Southeastern<br />

Publishing Corporation, pp. 12 13.<br />

Miller, M. G. & Johnson, A. D. (2002). White opaque<br />

switching in Candida albicans is controlled by matingtype<br />

locus homeodomain proteins and allows efficient<br />

mating. Cell, 110, 293 302.<br />

Miller, O. K. & Miller, H. H. (1988). Gasteromycetes.<br />

Morphological and Development Features with Keys <strong>to</strong> the<br />

Orders, Families, and Genera. Eureka, USA: Mad River<br />

Press.<br />

Miller, O. K., Henkel, T. W., James, T. Y. & Miller, S. L.<br />

(2001). Pseudotulos<strong>to</strong>ma, a remarkable new volvate<br />

genus in the Elaphomycetaceae from Guyana.<br />

Mycological Research, 105, 1268 1272.<br />

Miller, R. E. (1971). Evidence of sexuality in the<br />

cultivated mushroom. Mycologia, 63, 630 634.<br />

Miller, S. L. (1988). Early basidiospore formation in<br />

Lactarius lignyotellus. Mycologia, 80, 99 107.<br />

Mills, J. T. (1967). Spore dispersal and natural infection<br />

in the oat loose smut (Ustilago avenae). Transactions of<br />

the British Mycological Society, 50, 403 412.


770 REFERENCES<br />

Mims, C. W. (1982). Ultrastructure of the haus<strong>to</strong>rial<br />

apparatus of Exobasidium camelliae. Mycologia, 74,<br />

188 200.<br />

Mims, C. W. & Nickerson, N. L. (1986). Ultrastructure of<br />

the host pathogen relationship in red leaf disease<br />

of lowbush blueberry caused by the fungus<br />

Exobasidium vaccinii. Canadian Journal of Botany, 64,<br />

1338 1343.<br />

Mims, C. W. & Seabury, F. (1989). Ultrastructure of tube<br />

formation and basidiospore development in<br />

Ganoderma lucidum. Mycologia, 81, 754 764.<br />

Mims, C. W., Richardson, E. A. & Roberson, R. W. (1987).<br />

Ultrastructure of basidium and basidiospore development<br />

in three species of the fungus Exobasidium.<br />

Canadian Journal of Botany, 65, 1236 1244.<br />

Mims, C. W., Richardson, E. A. & Timberlake, W. E.<br />

(1988). Ultrastructural analysis of conidiophore<br />

development in the fungus Aspergillus nidulans using<br />

freeze-substitution. Pro<strong>to</strong>plasma, 144, 132 141.<br />

Minghetti, A. & Crespi-Perellino, N. (1999). The his<strong>to</strong>ry<br />

of ergot. In Ergot. The Genus Claviceps, ed. V. Křen & L.<br />

Cvak. Amsterdam: Harwood Academic Publishers,<br />

pp. 1 24.<br />

Minogue, K. P. & Fry, W. E. (1981). Effect of temperature,<br />

relative humidity, and rehydration rate on germination<br />

of dried sporangia of Phy<strong>to</strong>phthora infestans.<br />

Phy<strong>to</strong>pathology, 71, 1181 1184.<br />

Minter, D. W. (1984). New concepts in the interpretation<br />

of conidiogenesis in Deuteromycetes.<br />

Microbiological Sciences, 1, 86 89.<br />

Minter, D. W., Kirk, P. M. & Sut<strong>to</strong>n, B. C. (1982).<br />

Holoblastic phialides. Transactions of the British<br />

Mycological Society, 79, 75 93.<br />

Minter, D. W., Kirk, P. M. & Sut<strong>to</strong>n, B. C. (1983a). Thallic<br />

phialides. Transactions of the British Mycological Society,<br />

80, 39 66.<br />

Minter, D. W., Sut<strong>to</strong>n, B. C. & Brady, B. L. (1983b). What<br />

are phialides anyway? Transactions of the British<br />

Mycological Society, 81, 109 120.<br />

Minter, D. W., Cannon, P. F. & Peredo, H. L. (1987). South<br />

American species of Cyttaria (a remarkable and<br />

beautiful group of edible ascomycetes). Mycologist, 1,<br />

7 11.<br />

Mirza, J. H. & Cain, R. F. (1969). Revision of the<br />

genus Podospora. Canadian Journal of Botany, 47,<br />

1999 2048.<br />

Misra, J. K. (1998). Trichomycetes fungi associated<br />

with arthropods: review and world literature.<br />

Symbiosis, 24, 179 220.<br />

Misra, J. K. & Lichtwardt, R. W. (2000). Illustrated Genera<br />

of Trichomycetes. Fungal Symbionts of Insects and Other<br />

Arthropods. Enfield, USA: Science Publishers Inc.<br />

Mistry, A. (1977). Zygosporogenesis in Blakeslea trispora<br />

(Mucorales). Microbios, 20, 73 79.<br />

Mitani, S., Araki, S., Yamaguchi, T., et al. (2002).<br />

Biological properties of the novel fungicide cyazofamid<br />

against Phy<strong>to</strong>phthora infestans on <strong>to</strong>ma<strong>to</strong> and<br />

Pseudoperonospora cubensis on cucumber. Pest<br />

Management Science, 58, 139 145.<br />

Mitani, S., Sugimo<strong>to</strong>, K., Hayashi, H., et al. (2003). Effects<br />

of cyazofamid against Plasmodiophora brassicae<br />

Woronin on Chinese cabbage. Pest Management<br />

Science, 59, 287 293.<br />

Mitchell, R. T. & Deacon, J. W. (1986). Selective accumulation<br />

of zoospores of Chytridiomycetes and<br />

Oomycetes on cellulose and chitin. Transactions of the<br />

British Mycological Society, 86, 219 223.<br />

Mitchell, T. G. & Perfect, J. R. (1995). Cryp<strong>to</strong>coccosis in<br />

the era of AIDS 100 years after the discovery of<br />

Cryp<strong>to</strong>coccus neoformans. Clinical Microbiology Reviews, 8,<br />

515 548.<br />

Mithen, R. & Magrath, R. (1992). A contribution <strong>to</strong> the<br />

life his<strong>to</strong>ry of Plasmodiophora brassicae: secondary<br />

plasmodia development in root galls of Arabidopsis<br />

thaliana. Mycological Research, 96, 877 885.<br />

Mitra, M. (1931). A new bunt of wheat in India. Annals of<br />

Applied Biology, 18, 178 179.<br />

Mix, A. J. (1949). A monograph of the genus Taphrina.<br />

University of Kansas Science Bulletins, 30, 1 167.<br />

Miyake, H., Takemaru, T. & Ishikawa, T. (1980).<br />

Sequential production of enzymes and basidiospore<br />

formation in fruiting bodies of Coprinus macrorhizus.<br />

Archives of Microbiology, 126, 201 203.<br />

Moake, M. M., Padilla-Zakour, O. I. & Worobo, R. W.<br />

(2005). Comprehensive review of patulin control<br />

methods in foods. Comprehensive Reviews in Food Science<br />

and Food Safety, 1, 8 21.<br />

Molina, A., Hunt, M. D. & Ryals, J. A. (1998). Impaired<br />

fungicide activity in plants blocked in disease<br />

resistance signal transduction. Plant Cell, 10,<br />

1903 1914.<br />

Molina, M., Gil, C., Pla, J., Arroyo, J. & Nombrela, C.<br />

(2000). Protein localisation approaches for understanding<br />

yeast cell wall biogenesis. Microscopy<br />

Research and Technique, 51, 601 612.<br />

Molina, R. & Trappe, J. M. (1984). Mycorrhiza management<br />

in bareroot nurseries. In Forest Nursery Manual:<br />

Production of Bareroot Seedlings, ed. M. L. Duryea & T. D.<br />

Landis. Den Haag: Martinus Nijhoff and Dr. W. Junk<br />

Publishers, pp. 211 23.<br />

Molina, R. & Trappe, J. M. (1994). Biology of the<br />

ec<strong>to</strong>mycorrhizal genus, Rhizopogon. I. Host associations,<br />

host-specificity and pure culture syntheses.<br />

New Phy<strong>to</strong>logist, 126, 653 675.


REFERENCES<br />

771<br />

Molina, R., Trappe, J. M., Grubisha, L. C. & Spatafora,<br />

J. W. (1999). Rhizopogon. InEc<strong>to</strong>mycorrhizal <strong>Fungi</strong>: Key<br />

Genera in Profile, ed. J. W. G. Cairney & S. M. Chambers.<br />

Berlin: Springer Verlag, pp. 129 161.<br />

Moller, A. P. (1993). A fungus infecting domestic flies<br />

manipulates sexual behaviour of its host. Behavioural<br />

Ecology and Sociobiology, 33, 403 407.<br />

Mollicone, M. R. N. & Longcore, J. E. (1994). Zoospore<br />

ultrastructure of Monoblepharis polymorpha. Mycologia,<br />

86, 615 625.<br />

Mollicone, M. R. N. & Longcore, J. E. (1999). Zoospore<br />

ultrastructure of Gonapodya polymorpha. Mycologia, 91,<br />

727 734.<br />

Molnár, A., Sulyok, L. & Hornok, L. (1990). Parasexual<br />

recombination between vegetatively incompatible<br />

strains in Fusarium oxysporum. Mycological Research, 94,<br />

393 398.<br />

Momany, M. & Hamer, J. E. (1997). Relationship of actin,<br />

microtubules, and crosswall synthesis during septation<br />

in Aspergillus nidulans. Cell Motility and the<br />

Cy<strong>to</strong>skele<strong>to</strong>n, 38, 373 384.<br />

Momany, M., Richardson, E. A., van Sickle, C. & Jedd, G.<br />

(2002). Mapping Woronin body position in Aspergillus<br />

nidulans. Mycologia, 94, 260 266.<br />

Monaco, L. C. (1977). Consequences of the introduction<br />

of coffee rust in<strong>to</strong> Brazil. Annals of the New York<br />

Academy of Sciences, 287, 57 71.<br />

Moncalvo, J.-M., Lutzoni, F. M., Rehner, S. A., Johnson, J.<br />

& Vilgalys, R. (2000). Phylogenetic relationships of<br />

agaric fungi based on nuclear large subunit ribosomal<br />

DNA sequences. Systematic Biology, 49, 278 305.<br />

Moncalvo, J.-M., Vilgalys, R., Redhead, S. A., et al. (2002).<br />

One hundred and seventeen clades of euagarics.<br />

Molecular Phylogenetics and Evolution, 23, 357 400.<br />

Money, N. P. (1994). Osmotic adjustment and the role of<br />

turgor in mycelial fungi. In The Mycota I: Growth,<br />

Differentiation and Sexuality, ed. J. G. H. Wessels &<br />

F. Meinhardt. Berlin: Springer-Verlag, pp. 67 88.<br />

Money, N. P. (1997). Wishful thinking of turgor revisited:<br />

the mechanisms of fungal growth. Fungal<br />

Genetics and Biology, 21, 173 187.<br />

Money, N. P. (1998). More g’s than the space shuttle:<br />

ballis<strong>to</strong>spore discharge. Mycologia, 94, 547 558.<br />

Money, N. P. & Harold, F. M. (1992). Extension growth of<br />

the water mold Achlya: interplay of turgor and wall<br />

strength. Proceedings of the National Academy of Sciences<br />

of the USA, 89, 4245 4249.<br />

Money, N. P. & Harold, F. M. (1993). Two water molds<br />

can grow without measurable turgor pressure.<br />

Planta, 190, 426 430.<br />

Money, N. P. & Hill, T. W. (1997). Correlation between<br />

endoglucanase secretion and cell wall strength in<br />

oomycete hyphae: implications for growth and<br />

morphogenesis. Mycologia, 89, 777 785.<br />

Money, N. P. & Webster, J. (1985). Water stress and<br />

sporangial emptying in Achlya intricata. Botanical<br />

Journal of the Linnean Society, 91, 319 327.<br />

Money, N. P. & Webster, J. (1988). Cell wall permeability<br />

and its relationship <strong>to</strong> spore release in Achlya<br />

intricata. Experimental Mycology, 12, 169 179.<br />

Money, N. P. & Webster, J. (1989). Mechanism of<br />

sporangial emptying in Saprolegnia. Mycological<br />

Research, 92, 45 49.<br />

Money, N. P., Webster, J. & Ennos, R. (1988). Dynamics<br />

of sporangial emptying in Achlya intricata.<br />

Experimental Mycology, 12, 13 27.<br />

Money, N. P., That, T.-C. C.-T., Frederick, B. & Henson,<br />

J. M. (1998). Melanin synthesis is associated with<br />

changes in hyphopodial turgor, permeability, and<br />

wall rigidity in Gaeumannomyces graminis var. graminis.<br />

Fungal Genetics and Biology, 24, 240 251.<br />

Monschau, N., Stahmann, K.-P., Pielken, P. & Sahm, H.<br />

(1997). In vitro synthesis of b-(1-3)-glucan with a<br />

membrane fraction of Botrytis cinerea. Mycological<br />

Research, 101, 97 101.<br />

Montemurro, N. & Visconti, A. (1992). Alternaria<br />

metabolites chemical and biological data. In<br />

Alternaria Biology, Plant Disease and Metabolites, ed. J.<br />

Chelkowski & A. Visconti. Amsterdam: Elsevier,<br />

pp. 449 537.<br />

Montgomery, G. W. G. & Gooday, G. W. (1985).<br />

Phospholipid enzyme interactions of chitin<br />

synthase of Coprinus cinereus. FEMS Microbiology Letters,<br />

27, 29 33.<br />

Monzer, J. (1995). Actin-filaments are involved in<br />

cellular graviperception of the basidiomycete<br />

Flammulina velutipes. European Journal of Cell Biology,<br />

66, 151 156.<br />

Monzer, J. (1996). Cellular graviperception in the<br />

basidiomycete Flammulina velutipes can the nuclei<br />

serve as fungal sta<strong>to</strong>liths? European Journal of Cell<br />

Biology, 71, 216 220.<br />

Moore, D. (1994). Tissue formation. In The Growing<br />

Fungus, ed. N. A. R. Gow & G. M. Gadd. London:<br />

Chapman & Hall, pp. 423 465.<br />

Moore, D. (1998). Fungal Morphogenesis. Cambridge:<br />

Cambridge University Press.<br />

Moore, D. & Chiu, S. W. (2001). Fungal products as food.<br />

In Bio-Exploitation of Filamen<strong>to</strong>us <strong>Fungi</strong>, ed. S. B.<br />

Pointing & K. D. Hyde. Hong Kong: Fungal Diversity<br />

Press, pp. 223 251.<br />

Moore, D., Elhiti, M. M. Y. & Butler, R. D. (1979).<br />

Morphogenesis of the carpophore of Coprinus cinereus.<br />

New Phy<strong>to</strong>logist, 83, 695 722.


772 REFERENCES<br />

Moore, D., Hock, B., Greening, J. P., Kern, V. D., Frazer,<br />

L. N. & Monzer, J. (1996). Centenary Review.<br />

Gravimorphogenesis in agarics. Mycological Research,<br />

100, 257 273.<br />

Moore, E. D. & Miller, C. E. (1973). Resting body<br />

formation by rhizoidal fusion in Chytriomyces hyalinus.<br />

Mycologia, 65, 145 154.<br />

Moore, E. J. (1962). The on<strong>to</strong>geny of the sclerotia of<br />

Pyronema domesticum. Mycologia, 54, 312 316.<br />

Moore, E. J. (1963). The on<strong>to</strong>geny of the apothecia of<br />

Pyronema domesticum. American Journal of Botany, 50,<br />

37 44.<br />

Moore, E. J. & Korf, R. P. (1963). The genus Pyronema.<br />

Bulletin of the Torrey Botanical Club, 90, 33 42.<br />

Moore, N. Y., Pegg, K. G., Buddenhagen, I. W. & Bentley,<br />

S. (2001). Fusarium wilt of banana: a diverse clonal<br />

pathogen of a domesticated clonal host. In Fusarium.<br />

Paul E. Nelson Memorial Symposium, ed. B. A.<br />

Summerell, J. F. Leslie, D. Backhouse, W. L. Bryden &<br />

L. W. Burgess. St. Paul, USA: APS Press, pp. 212 224.<br />

Moore, R. T. (1955). Index <strong>to</strong> the Helicosporae. Mycologia,<br />

47, 90 103.<br />

Moore, R. T. (1985). The challenge of the dolipore/<br />

parenthesome septum. In Developmental Biology of<br />

Higher <strong>Fungi</strong>, ed. D. Moore, L. A. Cassel<strong>to</strong>n, D. A. Wood<br />

& J. C. Frankland. Cambridge: Cambridge University<br />

Press, pp. 175 212.<br />

Moore, R. T. (1987). The genera of Rhizoc<strong>to</strong>nia-like fungi:<br />

Ascorhizoc<strong>to</strong>nia, Cera<strong>to</strong>rhiza gen. nov., Epulorhiza gen.<br />

nov., Moniliopsis, and Rhizoc<strong>to</strong>nia. Mycotaxon, 29,<br />

91 99.<br />

Moore, R. T. (1990). The genus Lalaria gen. nov.:<br />

Taphrinales anamorphosum. Mycotaxon, 38,<br />

315 330.<br />

Moore, R. T. (1996). The dolipore/parenthesome in<br />

modern taxonomy. In Rhizoc<strong>to</strong>nia Species: Taxonomy,<br />

Molecular Biology, Ecology, Pathology and Disease Control,<br />

ed. S. Sneh, S. Jabayi-Hare, S. Neate & G. Dijst.<br />

Dordrecht, Netherlands: Kluwer, pp. 13 35.<br />

Moore, R. T. & Marchant, R. (1972). Ultrastructural<br />

characterization of the basidiomycete septum of<br />

Polyporus biennis. Canadian Journal of Botany, 50,<br />

2463 2469.<br />

Moore-Landecker, E. (1975). Effect of cultural conditions<br />

on apothecial morphogenesis in Pyronema<br />

domesticum. Canadian Journal of Botany, 53,<br />

2759 2769.<br />

Moore-Landecker, E. (1992). Physiology and biochemistry<br />

of ascocarp induction and development.<br />

Mycological Research, 96, 705 716.<br />

Morgan, W. M. (1983). Viability of Bremia lactucae<br />

oospores and stimulation of their germination by<br />

lettuce seedlings. Transactions of the British Mycological<br />

Society, 80, 403 408.<br />

Mori, Y., Sa<strong>to</strong>, Y. & Takamatsu, S. (2000). Evolutionary<br />

analysis of the powdery mildew fungi using nucleotide<br />

sequences of the nuclear ribosomal DNA.<br />

Mycologia, 92, 74 93.<br />

Morte, A., Lovisolo, C. & Schubert, A. (2000). Effect of<br />

drought stress on growth and water relations of the<br />

mycorrhizal association Helianthemum<br />

almeriense Terfezia claveryi. Mycorrhiza, 10, 115 119.<br />

Mor<strong>to</strong>n, C. O., Hirsch, P. R. & Kerry, B. R. (2004).<br />

Infection of plant-parasitic nema<strong>to</strong>des by nema<strong>to</strong>phagous<br />

fungi a review of the application of<br />

molecular biology <strong>to</strong> understand infection processes<br />

and <strong>to</strong> improve biological control. Nema<strong>to</strong>logy, 6,<br />

161 170.<br />

Mor<strong>to</strong>n, D. J. (1961). Tryptan blue and boiling lac<strong>to</strong>phenol<br />

for staining and clearing barley<br />

tissues infected with Ustilago nuda. Phy<strong>to</strong>pathology,<br />

51, 27 29.<br />

Mor<strong>to</strong>n, J. B. & Benny, G. L. (1990). Revised classification<br />

of arbuscular mycorrhizal fungi (Zygomycetes): a<br />

new order, Glomales, two new suborders, Glomineae<br />

and Gigasporineae, and two new families,<br />

Acaulosporaceae and Gigasporaceae, with an emendation<br />

of Glomaceae. Mycotaxon, 37, 471 491.<br />

Mor<strong>to</strong>n, J. B. & Bentivenga, S. P. (1994). Levels of<br />

diversity in endomycorrhizal fungi (Glomales,<br />

Zygomycetes) and their role in defining taxonomic<br />

and non-taxonomic groups. Plant and Soil, 159,<br />

47 59.<br />

Mor<strong>to</strong>n, J. B., Bentivenga, S. P. & Wheeler, W. W. (1993).<br />

Germ plasm in the International Collection of<br />

Arbuscular and Vesicular arbuscular Mycorrhizal<br />

<strong>Fungi</strong> (INVAM) and procedures for culture development,<br />

documentation and s<strong>to</strong>rage. Mycotaxon, 48,<br />

491 528.<br />

Moseman, J. G. (1966). Genetics of powdery mildews.<br />

Annual Review of Phy<strong>to</strong>pathology, 4, 269 291.<br />

Moseman, J. G. & Powers, H. R. (1957). Function and<br />

longevity of cleis<strong>to</strong>thecia of Erysiphe graminis f. sp.<br />

hordei. Phy<strong>to</strong>pathology, 47, 53 56.<br />

Moss, M. O. (1994). Biosynthesis of Aspergillus <strong>to</strong>xins<br />

non-afla<strong>to</strong>xins. In The Genus Aspergillus. From Taxonomy<br />

and Genetics <strong>to</strong> Industrial Application, ed. K. A. Powell,<br />

A. Renwick & J. F. Peberdy. New York: Plenum Press,<br />

pp. 29 50.<br />

Moss, M. O. (2002). Myco<strong>to</strong>xin review: 2. Fusarium.<br />

Mycologist, 16, 158 161.<br />

Moss, M. O. & Baker, T. (2002). Teaching techniques for<br />

mycology: 17. The pho<strong>to</strong>tropic response of Phycomyces<br />

blakesleeanus. Mycologist, 16, 23 26.


REFERENCES<br />

773<br />

Moss, M. O. & Long, M. T. (2002). Fate of patulin in the<br />

presence of the yeast Saccharomyces cerevisiae. Food<br />

Additives and Contaminants, 19, 387 399.<br />

Moss, S. T. (1975). Septal structure in the Trichomycetes<br />

with special reference <strong>to</strong> Astrep<strong>to</strong>nema gammari<br />

(Eccrinales). Transactions of the British Mycological<br />

Society, 65, 115 127.<br />

Moss, S. T. (1986). Biology and phylogeny of the<br />

Labyrinthulales and Thraus<strong>to</strong>chytriales. In The<br />

Biology of Marine <strong>Fungi</strong>, ed. S. T. Moss. Cambridge:<br />

Cambridge University Press, pp. 105 29.<br />

Moss, S. T. & Descals, E. (1986). A previously undescribed<br />

stage in the life cycle of Harpellales<br />

(Trichomycetes). Mycologia, 78, 213 222.<br />

Moss, S. T. & Lichtwardt, R. W. (1976). Development of<br />

trichospores and their appendages in Genistellospora<br />

homothallica and other Harpellales and finestructural<br />

evidence for the sporangial nature of<br />

trichospores. Canadian Journal of Botany, 54,<br />

2346 2364.<br />

Moss, S. T. & Lichtwardt, R. W. (1977). Zygospores of the<br />

Harpellales: an ultrastructural study. Canadian<br />

Journal of Botany, 55, 3099 3110.<br />

Moss, S. T. & Young, T. W. K. (1978). Phyletic considerations<br />

of the Harpellales and Asellariales<br />

(Trichomycetes, Zygomycotina) and the Kickxellales<br />

(Zygomycetes, Zygomycotina). Mycologia, 70,<br />

944 963.<br />

Mossebo, D. C. (1998). Étude de la polarité sexuelle chez<br />

Dacrymyces stillatus Nees:Fries (basidiomycète).<br />

Cryp<strong>to</strong>gamie Mycologie, 19, 285 300.<br />

Mossebo, D.-C. & Amougou, A. (2001). Mise en évidence<br />

de la dissolution des cloisons dolipores préalable à la<br />

migration nucléaire chez Dacrymyces stillatus<br />

Nees:Fries (Basidiomycete). Cryp<strong>to</strong>gamie Mycologie, 22,<br />

185 191.<br />

Mostert, L., Crous, P. W. & Petrini, O. (2000). Endophytic<br />

fungi associated with shoots and leaves of Vitis<br />

vinifera, with specific reference <strong>to</strong> the Phomopsis<br />

viticola complex. Sydowia, 52, 46 58.<br />

Mostert, L., Crous, P. W., Kang, J.-C. & Phillips, A. J. L.<br />

(2001). Species of Phomopsis and a Libertella sp.<br />

occurring on grapevines with specific reference <strong>to</strong><br />

South Africa: morphological, cultural, molecular<br />

and pathological characterization. Mycologia, 93,<br />

146 167.<br />

Motta, J. (1967). A note on the mi<strong>to</strong>tic apparatus in the<br />

rhizomorph meristem of Armillaria mellea. Mycologia,<br />

59, 370 375.<br />

Mouchacca, J. (1997). Thermophilic fungi: biodiversity<br />

and taxonomic status. Cryp<strong>to</strong>gamie Mycologie, 18,<br />

19 69.<br />

Mouchacca, J. (2000). Thermophilic fungi and applied<br />

research: a synopsis of name changes and synonymies.<br />

World Journal of Microbiology and Biotechnology,<br />

16, 881 888.<br />

Movahedi, S. & Heale, J. B. (1990). The roles of aspartic<br />

proteinase and endopectin lyase enzymes in the<br />

primary stages of infection and pathogenesis of<br />

various host tissues by different isolates of Botrytis<br />

cinerea Pers. ex Pers. Physiological and Molecular Plant<br />

Pathology, 36, 303 324.<br />

Mower, R. L. & Hancock, J. G. (1975a). Sugar composition<br />

of ergot honeydews. Canadian Journal of Botany,<br />

53, 2813 2825.<br />

Mower, R. L. & Hancock, J. G. (1975b). Mechanism of<br />

honeydew formation by Claviceps species. Canadian<br />

Journal of Botany, 53, 2826 2834.<br />

Moxham, S. E. & Buczacki, S. T. (1983). Chemical<br />

composition of the resting spore wall of<br />

Plasmodiophora brassicae. Transactions of the British<br />

Mycological Society, 80, 297 304.<br />

Moyersoen, B., Beever, R. E. & Martin, F. (2003). Genetic<br />

diversity of Pisolithus in New Zealand indicates<br />

multiple long-distance dispersal from Australia. New<br />

Phy<strong>to</strong>logist, 160, 569 579.<br />

Mtui, G. & Nakamura, Y. (2004). Lignin-degrading<br />

enzymes from mycelial cultures of basidiomycete<br />

fungi isolated in Tanzania. Journal of Chemical<br />

Engineering of Japan, 37, 113 118.<br />

Muehlstein, L. K., Porter, D. & Short, F. T. (1991).<br />

Labyrinthula zosterae sp. nov., the causative agent of<br />

wasting disease of eelgrass, Zostera marina. Mycologia,<br />

83, 180 191.<br />

Mueller, G. M. & Pine, E. M. (1994). DNA data provide<br />

evidence on the evolutionary relationships<br />

between mushrooms and false truffles. McIlvainea,<br />

11, 61 74.<br />

Mueller, W. C., Tessier, B. J. & Englander, L. (1986).<br />

Immunocy<strong>to</strong>chemical detection of fungi in the roots<br />

of Rhododendron. Canadian Journal of Botany, 64,<br />

718 723.<br />

Mugnier, J. & Mosse, B. (1987). Vesicular-arbuscular<br />

mycorrhizal infection in transformed root-inducing<br />

T-DNA roots grown axenically. Phy<strong>to</strong>pathology, 77,<br />

1045 1050.<br />

Mullens, B. A. & Rodriguez, J. L. (1985). Dynamics of<br />

En<strong>to</strong>mophthora muscae (En<strong>to</strong>mophthorales:<br />

En<strong>to</strong>mophthoraceae) conidial discharge from Musca<br />

domestica (Diptera: Muscidae) cadavers. Environmental<br />

En<strong>to</strong>mology, 14, 317 322.<br />

Muller, H. G. (1958). The constricting ring mechanism<br />

of two predacious Hyphomycetes. Transactions of the<br />

British Mycological Society, 41, 341 364.


774 REFERENCES<br />

Müller, E. & Tomasevic, M. (1957). Kulturversuche<br />

mit einigen Arten der Gattung Lep<strong>to</strong>sphaeria<br />

Ces. & de Not. Phy<strong>to</strong>pathologische Zeitschrift, 29,<br />

287 294.<br />

Müller, M. (1993). The hydrogenosome. Journal of General<br />

Microbiology, 139, 2879 2889.<br />

Müller, P. & Hilgenberg, W. (1986). Isomers of zeatin<br />

and zeatin riboside in clubroot tissue: evidence for<br />

trans-zeatin biosynthesis by Plasmodiophora brassicae.<br />

Physiologia Plantarum, 66, 245 250.<br />

Müller, W. H., Stalpers, J. A., van Aelst, A. C., van der<br />

Krift, T. P. & Boekhout, T. (1998a). Structural differences<br />

between the two types of basidiomycete septal<br />

pore caps. Microbiology, 144, 1721 1730.<br />

Müller, W. H., Stalpers, J. A., van Aelst, A. C., van der<br />

Krift, T. P. & Boekhout, T. (1998b). Field emission gunscanning<br />

electron microscopy of septal pore caps of<br />

selected species in the Rhizoc<strong>to</strong>nia s.l. complex.<br />

Mycologia, 90, 170 179.<br />

Müller, W. H., Koster, A. J., Humbel, B. M., et al. (2000).<br />

Au<strong>to</strong>mated electron <strong>to</strong>mography of the septal pore<br />

cap in Rhizoc<strong>to</strong>nia solani. Journal of Structural Biology,<br />

131, 10 18.<br />

Mullins, J. T. (1973). Lateral branch formation and<br />

cellulase production in the water molds. Mycologia,<br />

65, 1007 1014.<br />

Mullins, J. T. & Raper, J. R. (1965). Heterothallism in<br />

biflagellate aquatic fungi: preliminary genetic<br />

analysis. Science, 150, 1174 1175.<br />

Multani, D. S., Meeley, R. B., Paterson, A. H., et al. (1998).<br />

Plant-pathogen microevolution: molecular basis for<br />

the origin of a fungal disease in maize. Proceedings of<br />

the National Academy of Sciences of the USA, 95,<br />

1686 1691.<br />

Munn, A. L. (2000). The yeast endocytic membrane<br />

transport system. Microscopy Research and Technique,<br />

51, 547 562.<br />

Munn, E. A., Orpin, C. G. & Hall, F. J. (1981)<br />

Ultrastructural studies of the free zoospore of the<br />

rumen phycomycete Neocallimastix frontalis. Journal of<br />

General Microbiology, 125, 311 323.<br />

Murphy, J. W. (1996). Cell-mediated immunity. In The<br />

Mycota VI: Human and Animal Relationships, ed. D. H.<br />

Howard & J. D. Miller. Berlin: Springer-Verlag,<br />

pp. 67 97.<br />

Myers, R. B. & Cantino, E. C. (1974). The gamma particle.<br />

In Monographs in Developmental Biology 8, ed. A.<br />

Wolsky. New York: Karger, pp. 1 46.<br />

Myllys, L., Källersjö, M. & Tehler, A. (1998). A comparison<br />

of SSU rDNA data and morphological data in<br />

Arthoniales (Euascomycetes) phylogeny. Bryologist,<br />

101, 70 85.<br />

Myllys, L., Stenroos, S. & Thell, A. (2002). New genes for<br />

phylogenetic studies of lichenized fungi: glyceraldehyde-3-phosphate<br />

dehydrogenase and beta-tubulin<br />

genes. Lichenologist, 34, 237 246.<br />

Mylyk, O. M. (1976). Heteromorphism for heterokaryon<br />

incompatibility genes in natural populations of<br />

Neurospora crassa. Genetics, 83, 275 284.<br />

Nadeau, M. P., Dunphy, G. B. & Boisvert, J. L. (1995).<br />

Effects of physical fac<strong>to</strong>rs on the development of<br />

secondary conidia of Erynia conica (Zygomycetes:<br />

En<strong>to</strong>mophthorales), a pathogen of adult black flies<br />

(Diptera: Simuliidae). Experimental Mycology, 19,<br />

324 329.<br />

Nadeau, M. P., Dunphy, G. B. & Boisvert, J. L. (1996).<br />

Development of Erynia conica (Zygomycetes:<br />

En<strong>to</strong>mophthorales) on the cuticle of adult black flies<br />

Simulium rostratum and Simulium decorum (Diptera:<br />

Simuliidae). Journal of Invertebrate Pathology, 68,<br />

50 58.<br />

Nagahama, T., Hamamo<strong>to</strong>, M., Nakase, T., Takami, H. &<br />

Horikoshi, K. (2001). Distribution and identification<br />

of red yeasts in deep-sea environments around the<br />

northwest Pacific Ocean. An<strong>to</strong>nie van Leeuwenhoek, 80,<br />

101 110.<br />

Nagano, S. (2000). Modeling the model organism<br />

Dictyostelium discoideum. Development, Growth and<br />

Differentiation, 42, 541 550.<br />

Nagarajan, S. & Singh, D. V. (1990). Long-distance<br />

dispersal of rust pathogens. Annual Review of<br />

Phy<strong>to</strong>pathology, 28, 139 153.<br />

Nagarajan, S., Aujla, S. S., Nanda, G. S., et al. (1997).<br />

Karnal bunt (Tilletia indica) of wheat a review.<br />

Review of Plant Pathology, 76, 1207 1214.<br />

Nakagaki, T. (2001). Smart behavior of true slime mold<br />

in a labyrinth. Researches in Microbiology, 152,<br />

767 770.<br />

Nakagaki, T., Yamada, H. & Hara, M. (2004). Smart<br />

network solutions in an amoeboid organism.<br />

Biophysical Chemistry, 107, 1 5.<br />

Nakamura, A. & Kohama, K. (1999). Calcium regulation<br />

of the actin myosin interaction of Physarum polycephalum.<br />

International Review of Cy<strong>to</strong>logy, 191, 53 98.<br />

Nakase, T. (2000). Expanding world of ballis<strong>to</strong>sporous<br />

yeasts: distribution in the phyllosphere, systematics<br />

and phylogeny. Journal of General and Applied<br />

Microbiology, 46, 189 216.<br />

Nakatsuji, N. & Bell, E. (1980). Control by calcium of the<br />

contractility of Labyrinthula slimeways and of the<br />

translocation of Labyrinthula cells. Cell Motility, 1,<br />

17 19.<br />

Nannfeldt, J. A. (1932). Studien über die Morphologie<br />

und Systematik der nichtlichenisierten


REFERENCES<br />

775<br />

inoperculaten Discomyceten. Nova Acta Regiae Societas<br />

Scientiarum Upsaliensis IV, 8(2), 1 368.<br />

Nannfeldt, J. A. (1942). The Geoglossaceae of Sweden<br />

(with regard <strong>to</strong> surrounding countries). Arkiv för<br />

Botanik, 30A(4), 1 67.<br />

Nannfeldt, J. A. (1981). Exobasidium, a taxonomic reassessment<br />

applied <strong>to</strong> the European species. Symbolae<br />

Botanicae Uppsaliensis, 23, 1 72.<br />

Narisawa, K. & Hashiba, T. (1998). Development of<br />

resting spores on plants inoculated with a dikaryotic<br />

resting spore of Plasmodiophora brassicae. Mycological<br />

Research, 102, 949 952.<br />

Narisawa, K., Yamaoka, Y. & Katsuya, K. (1994). Mating<br />

type of isolates derived from the spermogonial state<br />

of Puccinia coronata var. coronata. Mycoscience, 35,<br />

131 135.<br />

Narisawa, K., Kageyama, K. & Hashiba, T. (1996).<br />

Efficient root infection with single resting spores of<br />

Plasmodiophora brassicae. Mycological Research, 100,<br />

855 858.<br />

Narisawa, K., Tokumasu, S. & Hashiba, T. (1998).<br />

Suppression of clubroot formation in Chinese<br />

cabbage by the root endophytic fungus, Heteroconium<br />

chae<strong>to</strong>spira. Plant Pathology, 47, 206 210.<br />

Nash, T. H., ed. (1996a). Lichen Biology. Cambridge:<br />

Cambridge University Press.<br />

Nash, T. H. (1996b). Nitrogen, its metabolism and<br />

potential contribution <strong>to</strong> ecosystems. In Lichen<br />

Biology, ed. T. H. Nash. Cambridge: Cambridge<br />

University Press, pp. 121 135.<br />

Nash, T. H. & Gries, C. (2002). Lichens as bioindica<strong>to</strong>rs of<br />

sulfur dioxide. Symbiosis, 33, 1 21.<br />

Nasmyth, K. (1983). Molecular analysis of a cell lineage.<br />

Nature, 302, 670 676.<br />

Nath, M. D., Sharma, S. L. & Kant, U. (2001). Growth of<br />

Albugo candida infected mustard callus in culture.<br />

Mycopathologia, 152, 147 153.<br />

Natvig, D. O. & Gleason, F. H. (1983). Oxygen uptake by<br />

obligately-fermentative aquatic fungi: absence of a<br />

cyanide-sensitive component. Archives of Microbiology,<br />

134, 5 8.<br />

Navarre, D. A. & Wolpert, T. J. (1999). Vic<strong>to</strong>rin induction<br />

of an apop<strong>to</strong>tic/senescence-like response in oats.<br />

Plant Cell, 11, 237 249.<br />

Nawawi, A. (1985). Basidiomycetes with branched<br />

water-borne conidia. Botanical Journal of the Linnean<br />

Society, 91, 51 60.<br />

Nawawi, A. & Webster, J. (1982). Sis<strong>to</strong>trema hamatum sp.<br />

nov., the teleomorph of Ingoldiella hamata. Transactions<br />

of the British Mycological Society, 78, 287 291.<br />

Neergaard, P. (1945). Danish species of Alternaria and<br />

Stemphylium. Copenhagen: E. Munksgaard.<br />

Neergaard, P. (1977). Seed Pathology, 2 volumes. London:<br />

Macmillan.<br />

Neiman, A. M. (2005). Ascospore formation in the yeast<br />

Saccharomyces cerevisiae. Microbiology and Molecular<br />

Biology Reviews, 69, 565 584.<br />

Nelson, A. C. & Backus, M. P. (1968). Ascocarp development<br />

in two homothallic Neurosporas. Mycologia, 60,<br />

16 28.<br />

Nelson, B. D. & Durán, R. (1984). Cy<strong>to</strong>logy and<br />

morphological development of basidia,<br />

dikaryons, and infective structures of<br />

Urocystis agropyri from wheat. Phy<strong>to</strong>pathology,<br />

74, 299 304.<br />

Nelson, J. C., Autrique, J. E., Fuentes-Dávila, G. &<br />

Sorrells, M. E. (1998). Chromosomal location of genes<br />

for resistance <strong>to</strong> Karnal bunt in wheat. Crop Science,<br />

38, 231 236.<br />

Nelson, P. E., Toussoun, T. A. & Cook, R. J., eds. (1981).<br />

Fusarium Diseases, Biology and Taxonomy. London, PA:<br />

Pennsylvania State University Press.<br />

Nelson, P. E., Toussoun, T. A. & Marasas, W. F. O. (1983).<br />

Fusarium Species. An Illustrated Manual for<br />

Identification. London, PA: Pennsylvania State<br />

University Press.<br />

Nelson, R. T., Yangco, B. G., Strake, D. T. & Cochrane, B. J.<br />

(1990). Genetic studies in the genus Basidiobolus.<br />

2. Phylogenetic relationships inferred from<br />

ribosomal DNA analysis. Experimental Mycology,<br />

14, 197 206.<br />

Németh, É. (1999). Parasitic production of ergot<br />

alkaloids. In Ergot. The Genus Claviceps, ed. V. Křen &<br />

L. Cvak. Amsterdam: Harwood Academic Publishers,<br />

pp. 303 19.<br />

Nes, W. D. (1990). Stereochemistry, sterol metamorphosis<br />

and evolution. In The Biochemistry, Structure and<br />

Utilization of Plant Lipids, ed. P. J. Quinn &<br />

J. L. Harwood. Portland: Portland Press, pp. 304 28.<br />

Nes, W. D. & Stafford, A. E. (1984). Side chain structural<br />

requirements for sterol-induced regulation<br />

of Phy<strong>to</strong>phthora cac<strong>to</strong>rum physiology. Lipids, 9,<br />

596 612.<br />

Nes, W. D., Patterson, G. W. & Bean, G. A. (1979). The<br />

effect of steroids and their solubilizing agents on<br />

mycelial growth of Phy<strong>to</strong>phthora cac<strong>to</strong>rum. Lipids, 14,<br />

458 462.<br />

Newhook, F. J., Young, B. R., Allen, S. D. & Allen, R. N.<br />

(1981). Zoospore motility of Phy<strong>to</strong>phthora cinnamomi<br />

in particulate substrates. Phy<strong>to</strong>pathologische Zeitschrift,<br />

101, 202 209.<br />

Newman, E. I. (1988). Mycorrhizal links between plants:<br />

their functioning and ecological significance.<br />

Advances in Ecological Research, 18, 243 270.


776 REFERENCES<br />

Newsham, K. K., Low, M. N. R., McLeod, A. R.,<br />

Greenslade, P. D. & Emmett, B. A. (1997). Ultraviolet-B<br />

radiation influences the abundance and distribution<br />

of phylloplane fungi on pedunculate oak (Quercus<br />

robur). New Phy<strong>to</strong>logist, 136, 287 297.<br />

Niccoli, T., Arellano, M. & Nurse, P. (2003). Role of Tea1p,<br />

Tea3p and Pom1p in the determination of cell ends<br />

in Schizosaccharomyces pombe. Yeast, 20, 1349 1358.<br />

Nickerson, A. W. & Raper, K. B. (1973). Macrocysts in the<br />

life cycle of the Dictyosteliaceae. II. Germination of<br />

the macrocysts. American Journal of Botany, 60,<br />

247 254.<br />

Niederhauser, J. S. (1991). Phy<strong>to</strong>phthora infestans: the<br />

Mexican connection. In Phy<strong>to</strong>phthora, ed. J. A. Lucas,<br />

R. C. Shat<strong>to</strong>ck, D. S. Shaw & L. R. Cooke. Cambridge:<br />

Cambridge University Press, pp. 25 45.<br />

Nielsen, K. A., Nicholson, R. A., Carver, T. L. W., Kunoh,<br />

H. & Oliver, R. P. (2000). First <strong>to</strong>uch: an immediate<br />

response <strong>to</strong> surface recognition in conidia of<br />

Blumeria graminis. Physiological and Molecular Plant<br />

Pathology, 56, 63 70.<br />

Nielsen, R. I. (1978). Sexual mutants of a heterothallic<br />

Mucor species, Mucor pusillus. Experimental Mycology, 2,<br />

193 197.<br />

Nielsen, J. (1988). Ustilago spp., smuts. Advances in Plant<br />

Pathology, 6, 483 490.<br />

Niemelä, T. & Korhonen, K. (1998). Taxonomy of the<br />

genus Heterobasidion. InHeterobasidion annosum.<br />

Biology, Ecology, Impact and Control, ed. S. Woodward, J.<br />

Stenlid, R. Karjalainan & A. Hüttermann.<br />

Wallingford, UK: CAB International, pp. 27 33.<br />

Nikoh, N. & Fukatsu, T. (2000). Interkingdom host<br />

jumping underground: phylogenetic analysis of<br />

en<strong>to</strong>mopathogenic fungi of the genus Cordyceps.<br />

Molecular Biology and Evolution, 17, 629 637.<br />

Niks, R. E., Walther, U., Jaiser, H., et al. (2000).<br />

Resistance against barley leaf rust (Puccinia hordei) in<br />

West-European spring barley germplasm. Agronomie,<br />

20, 769 782.<br />

Nishida, H. & Sugiyama, J. (1994). Archiascomycetes:<br />

detection of a major new lineage within the<br />

Ascomycota. Mycoscience, 35, 361 366.<br />

Noble, M. & Glynne, M. D. (1970). Wart disease of<br />

pota<strong>to</strong>es. F. A. O. Plant Protection Bulletin, 18, 125 135.<br />

Noble, R., Fermor, T. R., Lincoln, S., et al. (2003).<br />

Primordia initiation of mushroom (Agaricus bisporus)<br />

strains on axenic casing materials. Mycologia, 95,<br />

620 629.<br />

Nordbring-Hertz, B. (1972). Scanning electron microscopy<br />

of the nema<strong>to</strong>de-trapping organs in<br />

Arthrobotrys oligospora. Physiologia Plantarum, 26,<br />

279 284.<br />

Nordbring-Hertz, B. (1977). Nema<strong>to</strong>de-induced<br />

morphogenesis in the predacious fungus Arthrobotrys<br />

oligospora. Nema<strong>to</strong>logica, 23, 443 451.<br />

Nordbring-Hertz, B. (1988). Ecology and recognition in<br />

the nema<strong>to</strong>de nema<strong>to</strong>phagous fungus system.<br />

Advances in Microbial Ecology, 10, 81 114.<br />

Nordbring-Hertz, B. & Jansson, H.-B. (1984). Fungal<br />

development, predacity, and recognition of prey in<br />

nema<strong>to</strong>de-destroying fungi. In Current Perspectives in<br />

Microbial Ecology, ed. M. J. Klug & C. A. Reddy.<br />

Washing<strong>to</strong>n: American Society for Microbiology,<br />

pp. 327 333.<br />

Nordbring-Hertz, B., Jansson, H.-B., Persson, Y., Friman,<br />

E. & Dackman, C. (1995). Nema<strong>to</strong>phagous fungi. VHS<br />

film C1851. Göttingen: Institut für den<br />

Wissenschaftlichen Film.<br />

Normanly, J. (1997). Auxin metabolism. Physiologia<br />

Plantarum, 100, 431 442.<br />

Nosanchuk, J. D. (2005). Protective antibodies and<br />

endemic dimorphic fungi. Current Molecular Medicine,<br />

5, 435 442.<br />

Nouhra, E. R., Hor<strong>to</strong>n, T. R., Cazares, E. & Castellano, M.<br />

(2005). Morphological and molecular characterization<br />

of selected Ramaria mycorrhizae. Mycorrhiza, 15,<br />

55 59.<br />

Nout, M. J. R. & Aidoo, K. E. (2002). Asian fungal<br />

fermented food. In The Mycota X: Industrial<br />

Applications, ed. H. D. Osiewacz. Berlin: Springer-<br />

Verlag, pp. 23 47.<br />

Nurse, P. (1990). Universal control mechanism regulating<br />

onset of M-phase. Nature, 344, 503 508.<br />

Nurse, P. (2002). Cyclin dependent kinases and cell cycle<br />

control (Nobel lecture). Chembiochem, 3, 596 603.<br />

Nuss, D. L. (1992). Biological control of chestnut blight:<br />

an example of virus-mediated attenuation of fungal<br />

pathogenesis. Microbiological Reviews, 56, 561 576.<br />

Nuss, D. L. (1996). Using hypoviruses <strong>to</strong> probe and<br />

perturb signal transduction processes underlying<br />

fungal pathogenesis. Plant Cell, 8, 1845 1853.<br />

Nuss, I., Jennings, D. H. & Veltkamp, C. J. (1991).<br />

Morphology of Serpula lacrymans. InSerpula lacrymans:<br />

Fundamental Biology and Control Strategies, ed. D. H.<br />

Jennings & A. F. Bravery. Chichester: Wiley, pp. 9 38.<br />

Nutting, R., Rapoport, H. & Machlis, L. (1968). The<br />

structure of sirenin. Journal of the American Chemical<br />

Society, 90, 6434 6438.<br />

Nyhlén, L. & Unestam, T. (1980). Wound reactions and<br />

Aphanomyces astaci growth in crayfish cuticle. Journal<br />

of Invertebrate Pathology, 36, 187 197.<br />

Nylander, W. (1866). Les lichens du Jardin du<br />

Luxembourg. Bulletin de la Société Botanique de France,<br />

Lettres Botaniques, 13, 364 372.


REFERENCES<br />

777<br />

Obermayer, W. & Poelt, J. (1992). Contributions <strong>to</strong> the<br />

knowledge of the lichen flora of the Himalayas. III. On<br />

Lecanora somervellii Paulson (lichenized Ascomycotina,<br />

Lecanoraceae). Lichenologist, 24, 111 117.<br />

Oberwinkler, F. (1993). Genera in a monophyletic<br />

group: the Dacrymycetales. Mycologia Helvetica, 6,<br />

35 72.<br />

O’Brien, E. & Dietrich, D. R. (2005). Ochra<strong>to</strong>xin A: the<br />

continuing enigma. Critical Reviews in Toxicology, 35,<br />

33 60.<br />

O’Connell, R. J., Bailey, J. A. & Richmond, D. V. (1985).<br />

Cy<strong>to</strong>logy and physiology of infection of Phaseolus<br />

vulgaris infected by Colle<strong>to</strong>trichum lindemuthianum.<br />

Physiological Plant Pathology, 27, 75 98.<br />

O’Connell, R. J., Perfect, S., Hughes, B., et al. (2000).<br />

Dissecting the cell biology of Colle<strong>to</strong>trichum infection<br />

processes. In Colle<strong>to</strong>trichum. Host Specificity, Pathology,<br />

and Host Pathogen Interaction, ed. D. Prusky, S.<br />

Freeman & M. B. Dickman. St. Paul, USA: APS Press,<br />

pp. 57 77.<br />

Oda, T., Tanaka, C. & Tsuda, M. (2004). Molecular<br />

phylogeny and biogeography of the widely distributed<br />

Amanita species, A. muscaria and A. pantherina.<br />

Mycological Research, 108, 885 896.<br />

Odds, F. C. (1994). Candida species and virulence. ASM<br />

News, 60, 313 318.<br />

Odds, F. C. (2003). Coccidioidomycosis: flying conidia<br />

and severed heads. Mycologist, 17, 37 40.<br />

Odom, A., Muir, S., Lim, E., et al. (1997). Calcineurin is<br />

required for virulence of Cryp<strong>to</strong>coccus neoformans.<br />

EMBO Journal, 16, 2576 2589.<br />

O’Donnell, K. (1996). Progress <strong>to</strong>wards a phylogenetic<br />

classification of Fusarium. Sydowia, 48, 57 70.<br />

O’Donnell, K., Kistler, H. C., Cigelnik, E. & Ploetz, R. C.<br />

(1998). Multiple evolutionary origins of the fungus<br />

causing Panama disease of banana: concordant<br />

evidence from nuclear and mi<strong>to</strong>chondrial gene<br />

genealogies. Proceedings of the National Academy of<br />

Sciences of the USA, 95, 2044 2049.<br />

O’Donnell, K. L. (1979). Zygomycetes in Culture. Athens,<br />

USA: Department of Botany, University of Georgia.<br />

O’Donnell, K. L., Fields, W. G. & Hooper, G. R. (1974).<br />

Scanning ultrastructural on<strong>to</strong>geny of cleis<strong>to</strong>hymenial<br />

apothecia of the operculate Discomycete<br />

Ascobolus furfuraceus. Canadian Journal of Botany, 52,<br />

1653 1656.<br />

O’Donnell, K. L., Hooper, G. R. & Fields, W. G. (1976).<br />

Zygosporogenesis in Phycomyces blakesleeanus.<br />

Canadian Journal of Botany, 54, 2573 2586.<br />

O’Donnell, K. L., Ellis, J. J., Hesseltine, C. W. & Hooper,<br />

G. R. (1977a). Zygosporogenesis in Gilbertella persicaria.<br />

Canadian Journal of Botany, 55, 662 675.<br />

O’Donnell, K. L., Ellis, J. J., Hesseltine, C. W. & Hooper,<br />

G. R. (1977b). Azygosporogenesis in Mucor azygosporus.<br />

Canadian Journal of Botany, 55, 2712 2720.<br />

O’Donnell, K. L., Ellis, J. J., Hesseltine, C. W. & Hooper,<br />

G. R. (1977c). Morphogenesis of azygospores induced<br />

in Gilbertella persicaria (þ) by imperfect hybridization<br />

with Rhizopus s<strong>to</strong>lonifer. Canadian Journal of Botany, 55,<br />

2721 2727.<br />

O’Donnell, K. L., Flegler, S. L., Ellis, J. J. & Hesseltine,<br />

C. W. (1978a). The Zygorhynchus zygosporangium and<br />

zygospore. Canadian Journal of Botany, 56, 1061 1073.<br />

O’Donnell, K. L., Flegler, S. L. & Hooper, G. R. (1978b).<br />

Zygosporangium and zygospore formation in<br />

Phycomyces nitens. Canadian Journal of Botany, 56,<br />

91 100.<br />

O’Donnell, K. L., Cigelnik, E., Weber, N. S. & Trappe, J. M.<br />

(1997). Phylogenetic relationships among ascomyce<strong>to</strong>us<br />

truffles and the true and false morels inferred<br />

from 18S and 28S ribosomal DNA sequence analysis.<br />

Mycologia, 87, 48 65.<br />

O’Donnell, K. L., Cigelnik, E. & Benny, G. L. (1998).<br />

Phylogenetic relationships among the Harpellales<br />

and Kickxellales. Mycologia, 90, 624 639.<br />

O’Donnell, K. L., Lutzoni, F. M., Ward, T. J. & Benny, G. L.<br />

(2001). Evolutionary relationships among<br />

Mucoralean fungi (Zygomycota). Evidence for family<br />

polyphyly on a large scale. Mycologia, 93, 286 296.<br />

Oeser, B., Tenberge, K. B., Moore, S. M., et al. (2002).<br />

Pathogenic development of Claviceps purpurea. In<br />

Molecular Biology of Fungal Development, ed. H. D.<br />

Osiewacz. New York: Marcel Dekker Inc., pp. 419 455.<br />

Ogawa, H. & Sugiyama, J. (2000). Evolutionary relationships<br />

of the cleis<strong>to</strong>thecial genera with<br />

Penicillium, Geosmithia, Merimbla and Sarophorum<br />

anamorphs as inferred from 18S rDNA sequence<br />

divergence. In Integration of Modern Taxonomic Methods<br />

for Penicillium and Aspergillus Classification, ed. R. A.<br />

Samson & J. I. Pitt. Amsterdam: Harwood Academic<br />

Publishers, pp. 149 61.<br />

Ogawa, H., Yoshimura, A. & Sugiyama, J. (1997).<br />

Phylogenetic origins of the anamorphic genus<br />

Geosmithia and the relationships of the cleis<strong>to</strong>thecial<br />

genera: evidence from 18S, 5S and 28S sequence<br />

analyses. Mycologia, 89, 756 771.<br />

Ohtake, M., Yamamo<strong>to</strong>, H. & Uchiyama, T. (1999)<br />

Influences of metabolic inhibi<strong>to</strong>rs and hydrolytic<br />

enzymes on the adhesion of appressoria of Pyricularia<br />

oryzae <strong>to</strong> wax-coated cover-glasses. Bioscience,<br />

Biotechnology, and Biochemistry, 63, 978 982.<br />

Ojha, M. & Turian, G. (1971). Interspecific transformation<br />

and DNA characteristics in Allomyces. Molecular<br />

and General Genetics, 112, 49 59.


778 REFERENCES<br />

Okada, G., Jacobs, K., Kirisits, T., et al. (2000).<br />

Epitypification of Graphium penicillioides Corda, with<br />

comments on the phylogeny and taxonomy of<br />

graphium-like synnema<strong>to</strong>us fungi. Studies in<br />

Mycology, 45, 169 188.<br />

Okafor, J. I., Testrake, D., Mushinsky, H. R. & Yangco,<br />

B. G. (1984). A Basidiobolus sp. and its association with<br />

reptiles and amphibians in southern Florida.<br />

Sabouraudia, 22, 47 51.<br />

Okamo<strong>to</strong>, K. & Shaw, J. M. (2005). Mi<strong>to</strong>chondrial<br />

morphology and dynamics in yeast and multicellular<br />

eukaryotes. Annual Review of Genetics, 39, 503 536.<br />

Olive, L. S. (1947). Notes on the Tremellales of Georgia.<br />

Mycologia, 39, 90 108.<br />

Olive, L. S. (1967). The Pro<strong>to</strong>stelida a new order of the<br />

Myce<strong>to</strong>zoa. Mycologia, 59, 1 29.<br />

Olive, L. S. (1975). The Myce<strong>to</strong>zoans. New York: Academic<br />

Press.<br />

Oliver, S. G. (1991). ‘Classical’ yeast biotechnology. In<br />

Saccharomyces, ed. M. F. Tuite & S. G. Oliver. London:<br />

Plenum Press, pp. 213 48.<br />

Olivier, C., Berbee, M. L., Shoemaker, R. A. & Loria, R.<br />

(2000). Molecular phylogenetic support from ribosomal<br />

DNA sequences for origin of Helminthosporium<br />

from Lep<strong>to</strong>sphaeria-like loculoascomycete ances<strong>to</strong>rs.<br />

Mycologia, 92, 736 746.<br />

Olsen, J. H., Dragsted, L. & Autrup, H. (1988). Cancer risk<br />

and occupational exposure <strong>to</strong> afla<strong>to</strong>xins in<br />

Denmark. British Journal of Cancer, 58, 392 6<br />

Olson, L. W. (1974). Meiosis in the aquatic Phycomycete<br />

Allomyces macrogynus. Comptes Rendus des Travaux du<br />

Labora<strong>to</strong>ire Carlsberg, 40, 113 124.<br />

Olson, L. W. (1984). Allomyces a different fungus. Opera<br />

Botanica, 73, 1 96.<br />

Olson, L. W. & Fuller, M. S. (1968). Ultrastructural<br />

evidence for the biflagellate origin of the uniflagellate<br />

fungal zoospore. Archiv für Mikrobiologie, 62,<br />

237 250.<br />

Olson, L. W. & Reichle, R. (1978). Synap<strong>to</strong>nemal<br />

complex formation and meiosis in the resting<br />

sporangium of Blas<strong>to</strong>cladiella emersonii. Pro<strong>to</strong>plasma,<br />

97, 261 273.<br />

O’Neal, M. A. & Schoenenberger, K. R. (2003). A<br />

Rhizocarpon geographicum growth curve for the<br />

cascade range of Washing<strong>to</strong>n and northern Oregon,<br />

USA. Quaternary Research, 60, 233 241.<br />

Ono, Y. (2002). The diversity of nuclear cycle in<br />

microcyclic rust fungi (Uredinales) and its ecological<br />

and evolutionary implications. Mycoscience, 43,<br />

421 439.<br />

Orlowski, M. (1991). Mucor dimorphism. Microbiological<br />

Reviews, 55, 234 258.<br />

Orlowski, M. (1995). Gene expression in Mucor<br />

dimorphism. Canadian Journal of Botany, 73,<br />

S326 S334.<br />

Ormrod, D. J., O’Reilly, H. J., van der Kamp, B. J. & Borno,<br />

C. (1984). Epidemiology, cultivar susceptibility, and<br />

chemical control of Gymnosporangium fuscum in<br />

British Columbia. Canadian Journal of Plant Pathology,<br />

6, 63 70.<br />

Orpin, C. G. (1974). The rumen flagellate Callimastix<br />

frontalis: does sequestration occur? Journal of General<br />

Microbiology, 84, 395 398.<br />

Orpin, C. G. & Munn, E. A. (1986). Neocallimastix patriciarum<br />

sp. nov., a new member of the<br />

Neocallimasticaceae inhabiting the rumen of sheep.<br />

Transactions of the British Mycological Society, 86,<br />

178 181.<br />

Ortega, J. K. E. (1990). Governing equations for plant<br />

cell growth. Physiologia Plantarum, 79, 116 121.<br />

Ortega, J. K. E., Lesh-Laurie, G. E., Espinosa, M. A., et al.<br />

(2003). Helical growth of stage-IVb sporangiophores<br />

of Phycomyces blakesleeanus: the relationship between<br />

rotation and elongation growth rates. Planta, 216,<br />

716 722.<br />

Ortiz-García, S., Gernandt, D. S., S<strong>to</strong>ne, J. K., et al. (2003).<br />

Phylogenetics of Lophodermium from pine. Mycologia,<br />

95, 846 859.<br />

Or<strong>to</strong>n, P. D. & Watling, R. (1979). Coprinaceae Part 1:<br />

Coprinus. British Fungus Flora 2. Edinburgh: Royal<br />

Botanic Garden.<br />

Osiewacz, H. D. (2002). Aging in fungi: role of mi<strong>to</strong>chondria<br />

in Podospora anserina. Mechanisms of Ageing<br />

and Development, 123, 755 764.<br />

Osiewacz, H. D. & Kimpel, E. (1999).<br />

Mi<strong>to</strong>chondrial nuclear interactions and lifespan<br />

control in fungi. Experimental Geron<strong>to</strong>logy, 34,<br />

901 909.<br />

Oso, B. (1969). Electron microscopy of ascus development<br />

in Ascobolus. Annals of Botany N.S., 33, 205 209.<br />

Ota, Y., Fukuda, K. & Suzuki, K. (1998). The nonheterothallic<br />

life cycle of Japanese Armillaria mellea.<br />

Mycologia, 90, 396 405.<br />

Otani, H. & Kohmo<strong>to</strong>, K. (1992). Host-specific <strong>to</strong>xins of<br />

Alternaria species. In Alternaria, Biology, Plant Diseases<br />

and Metabolites, ed. J. Chelkowski & A. Visconti.<br />

Amsterdam: Elsevier, pp. 123 56.<br />

Ott, S. (1987). Sexual reproduction and developmental<br />

adaptations in Xanthoria parietina. Nordic Journal of<br />

Botany, 7, 219 228.<br />

Ott, S. & Lumbsch, H. T. (2001). Morphology and<br />

phylogeny of ascomycete lichens. In The Mycota IX:<br />

Fungal Associations, ed. B. Hock. Berlin: Springer-<br />

Verlag, pp. 189 210.


REFERENCES<br />

779<br />

Ott, S., Meier, T. & Jahns, H. M. (1995). Development,<br />

regeneration, and parasitic interactions between the<br />

lichens Fulgensia bracteata and Toninia caeruleonigricans.<br />

Canadian Journal of Botany, 73, S595 S602.<br />

Ouimette, D. G. & Coffey, M. D. (1990). Symplastic entry<br />

and phloem translocation of phosphonate. Pesticide<br />

Biochemistry and Physiology, 38, 18 25.<br />

Ower, R. (1982). Notes on the development of the morel<br />

ascocarp: Morchella esculenta. Mycologia, 74, 142 144.<br />

Oxenbøll, K. (1994). Aspergillus enzymes and industrial<br />

uses. In The Genus Aspergillus. From Taxonomy and<br />

Genetics <strong>to</strong> Industrial Application, ed. K. A. Powell, A.<br />

Renwick & J. F. Peberdy. New York: Plenum Press,<br />

pp. 147 154.<br />

Oxford, A. E., Raistrick, H. & Simonart, P. (1939). Studies<br />

on the biochemistry of microorganisms. 60.<br />

Griseofulvin, a metabolic product of Penicillium<br />

griseofulvum Dierckx. Biochemical Journal, 33, 240 248.<br />

Pacioni, G. (1991). Effect of Tuber metabolites on the<br />

rhizosphere environment. Mycological Research, 95,<br />

1355 1358.<br />

Pacioni, G., Bellina-Agostinone, C. & D’An<strong>to</strong>nio, M.<br />

(1990). Odour composition of the Tuber melanosporum<br />

complex. Mycological Research, 94, 201 204.<br />

Pady, S. M. & Johns<strong>to</strong>n, C. O. (1955). The concentration<br />

of airborne rust spores in relation <strong>to</strong> epidemiology<br />

of wheat rusts in Kansas in 1954. Plant Disease<br />

Reporter, 39, 463 466.<br />

Page, R. M. (1959). Stimulation of asexual reproduction<br />

of Pilobolus by Mucor plumbeus. American Journal of<br />

Botany, 46, 579 585.<br />

Page, R. M. (1960). The effect of ammonia on growth<br />

and reproduction of Pilobolus kleinii. Mycologia, 52,<br />

480 489.<br />

Page, R. M. (1964). Sporangium discharge in Pilobolus: a<br />

pho<strong>to</strong>graphic study. Science, 146, 925 927.<br />

Page, R. M. & Curry, G. M. (1966). Studies on pho<strong>to</strong>tropism<br />

of young sporangiophores of Pilobolus kleinii.<br />

Pho<strong>to</strong>chemistry and Pho<strong>to</strong>biology, 5, 31 40.<br />

Page, R. M. & Kennedy, D. (1964). Studies on the velocity<br />

of discharged sporangia of Pilobolus kleinii. Mycologia,<br />

56, 363 368.<br />

Page, R. M. & Humber, R. A. (1973). Pho<strong>to</strong>tropism in<br />

Conidiobolus coronatus. Mycologia, 65, 335 354.<br />

Paktitis, S., Grant, B. & Lawrie, A. (1986). Surface<br />

changes in Phy<strong>to</strong>phthora palmivora zoospores following<br />

induced differentiation. Pro<strong>to</strong>plasma, 135,<br />

119 129.<br />

Palecek, S. P., Parikh, A. S. & Kron, S. J. (2002). Sensing,<br />

signalling and integrating physical processes during<br />

Saccharomyces cerevisiae invasive and filamen<strong>to</strong>us<br />

growth. Microbiology, 148, 893 907.<br />

Palfreyman, J. W. & White, N. A. (2003). Everything you<br />

wanted <strong>to</strong> know about the dry-rot fungus but were<br />

afraid <strong>to</strong> ask. Microbiology Today, 30, 107 109.<br />

Palfreyman, J. W., White, N. A., Buultjens, T. E. J. &<br />

Glancy, H. (1995). The impact of current research on<br />

the treatment of infestations by the dry rot fungus<br />

Serpula lacrymans. International Biodeterioration and<br />

Biodegradation, 35, 369 395.<br />

Palm, M. E. (1999). Mycology and world trade: a view<br />

from the front line. Mycologia, 91, 1 12.<br />

Palmer, C.-L. & Skinner, W. (2002). Mycosphaerella<br />

graminicola: latent infection, crop devastation and<br />

genomics. Molecular Plant Pathology, 3, 63 70.<br />

Panabières, F., Ponchet, M., Allasia, V., Cardin, L. &<br />

Ricci, P. (1997). Characterization of border species<br />

among Pythiaceae: several Pythium isolates produce<br />

elicitins, typical proteins of Phy<strong>to</strong>phthora spp.<br />

Mycological Research, 101, 1459 1468.<br />

Pande, S., Siddique, K. H. M., Kishore, G. K., et al. (2005).<br />

Ascochyta blight of chickpea (Cicer arietinum L.): a<br />

review of biology, pathogenicity, and disease<br />

management. Australian Journal of Agricultural<br />

Research, 56, 317 332.<br />

Papavizas, G. C. (1985). Trichoderma and Gliocladium:<br />

biology, ecology and potential for biocontrol. Annual<br />

Review of Phy<strong>to</strong>pathology, 23, 23 54.<br />

Papavizas, G. C. & Ayers, W. A. (1974). Aphanomyces<br />

species and their root rot diseases in pea and<br />

sugarbeet. US Department of Agriculture Technical<br />

Bulletin, 1485, 1 158.<br />

Papierok, B. & Hajek, A. E. (1997). <strong>Fungi</strong>:<br />

En<strong>to</strong>mophthorales. In Manual of Techniques in Insect<br />

Pathology, ed. L. A. Lacey. London: Academic Press,<br />

pp. 187 212.<br />

Parbery, D. G. (1996a). Spermatial states of fungi are<br />

andromorphs. Mycological Research, 100, 1400.<br />

Parbery, D. G. (1996b). Trophism and the<br />

association of fungi with plants. Biological Reviews,<br />

71, 473 527.<br />

Parguey-Leduc, A. & Janex-Favre, M. C. (1982). La paroi<br />

des asques chez les Pyrenomycètes. Étude ultrastructurale.<br />

I. Les asques bituniqués typiques.<br />

Canadian Journal of Botany, 60, 1222 1230.<br />

Parguey-Leduc, A. & Janex-Favre, M. C. (1984). La paroi<br />

des asques chez les Pyrenomycètes. Étude ultrastructurale.<br />

II. Les asques unituniqués. Cryp<strong>to</strong>gamie<br />

Mycologie, 5, 171 187.<br />

Parguey-Leduc, A., Janex-Favre, M. C. & Montant, C.<br />

(1991). L’ascocarpe du Tuber melanosporum Vitt.<br />

(Truffe noire du Périgord. Discomycètes): structure<br />

de la glèbe. II. Les veines steriles. Cryp<strong>to</strong>gamie<br />

Mycologie, 12, 165 182.


780 REFERENCES<br />

Park, D. (1974) Aquatic hyphomycetes in non-aquatic<br />

habitats. Transactions of the British Mycological Society,<br />

63, 179 183.<br />

Park, R. F., Burdon, J. J. & McIn<strong>to</strong>sh, R. A. (1995). Studies<br />

on the origin, spread, and evolution of an important<br />

group of Puccinia recondita f. sp. tritici pathotypes in<br />

Australasia. European Journal of Plant Pathology, 101,<br />

613 622.<br />

Park, R. F., Burdon, J. J. & Jahoor, A. (1999). Evidence for<br />

somatic hybridization in nature in Puccinia recondita<br />

f. sp. tritici, the leaf rust pathogen of wheat.<br />

Mycological Research, 103, 715 723.<br />

Parladé, J., Pera, J. & Alvarez, I. F. (1996). Inoculation of<br />

containerized Pseudotsuga menziesii and Pinus pinaster<br />

seedlings with spores of five species of ec<strong>to</strong>mycorrhizal<br />

fungi. Mycorrhiza, 6, 237 245.<br />

Parry, D. W. (1990). Plant Pathology in Agriculture.<br />

Cambridge: Cambridge University Press.<br />

Partida-Martinez, L. P. & Hertweck, C. (2005).<br />

Pathogenic fungus harbours endosymbiotic bacteria<br />

for <strong>to</strong>xin production. Nature, 437, 884 888.<br />

Pataky, J. K. & Chandler, M. A. (2003). Production of<br />

huitlacoche, Ustilago maydis: timing inoculation and<br />

controlling pollination. Mycologia, 95, 1261 1270.<br />

Pa<strong>to</strong>uillard, N. T. (1887). Les Hyménomycètes d’Europe.<br />

Ana<strong>to</strong>mie Générale et Classification des Champignons<br />

Supérieurs. Paris: Librairie Paul Klincksieck.<br />

Paul, D., Mukhopadhyay, R., Chatterjee, B. P. & Guha,<br />

A. K. (2002). Nutritional profile of food yeast<br />

Kluyveromyces fragilis biomass grown on whey. Applied<br />

Biochemistry and Biotechnology, 97, 209 218.<br />

Paulin-Mahady, A. E., Harring<strong>to</strong>n, T. C. & McNew, D.<br />

(2002). Phylogenetic and taxonomic evaluation of<br />

Chalara, Chalaropsis, and Thielaviopsis anamorphs<br />

associated with Cera<strong>to</strong>cystis. Mycologia, 94, 62 72.<br />

Pažou<strong>to</strong>vá, S., Olšovská, J., Linka, M., Kolínská, R.&<br />

Flieger, M. (2000). Chemoraces and habitat specialisation<br />

of Claviceps purpurea populations. Applied and<br />

Environmental Microbiology, 66, 5419 5425.<br />

Pazzagli, L., Cappugi, G., Manao, G., et al. (1999).<br />

Purification, characterization, and amino acid<br />

sequence of cera<strong>to</strong>-platanin, a new phy<strong>to</strong><strong>to</strong>xic<br />

protein from Cera<strong>to</strong>cystis fimbriata f. sp. platani.<br />

Journal of Biological Chemistry, 274, 24 959 24 964.<br />

Pearce, D. A. (1998). PCR as a <strong>to</strong>ol for the investigation<br />

of seed-borne disease. In Applications of PCR in<br />

Mycology, ed. P. D. Bridge, D. K. Arora, C. A. Reddy &<br />

R. P. Elander. Wallingford, UK: CAB International,<br />

pp. 309 24.<br />

Peberdy, J. F. (1994). Protein secretion in filamen<strong>to</strong>us<br />

fungi trying <strong>to</strong> understand a highly productive<br />

black box. Trends in Biotechnology, 12, 50 57.<br />

Pedras, M. S. C. & Biesenthal, C. J. (1998). Production of<br />

the host-selective phy<strong>to</strong><strong>to</strong>xin phomalide by isolates<br />

of Lep<strong>to</strong>sphaeria maculans and its correlation with<br />

sirodesmin PL production. Canadian Journal of<br />

Microbiology, 44, 547 553.<br />

Pedras, M. S. C., Okanga, F. I., Zaharia, I. L. & Khan, A. Q.<br />

(2000). Phy<strong>to</strong>alexins from crucifers: synthesis,<br />

biosynthesis, and biotransformation. Phy<strong>to</strong>chemistry,<br />

53, 161 176.<br />

Pegg, G. F. (1985). Presidential address. Life in a black<br />

hole the micro-environment of the vascular<br />

pathogen. Transactions of the British Mycological Society,<br />

85, 1 20.<br />

Pegler, D. N. (1991). Taxonomy, identification and<br />

recognition of Serpula lacrymans. In Serpula lacrymans.<br />

Fundamental Biology and Control Strategies, ed. D. H.<br />

Jennings & A. F. Bravery. Chichester: John Wiley &<br />

Sons, pp. 1 7.<br />

Pegler, D. N. (1996). Hyphal analysis of basidiomata.<br />

Mycological Research, 100, 129 142.<br />

Pegler, D. N. (2000). Taxonomy, nomenclature and<br />

description of Armillaria. In Armillaria Root Rot: Biology<br />

and Control of Honey Fungus, ed. R. T. V. Fox. Andover,<br />

UK: Intercept Ltd, pp. 81 93.<br />

Pegler, D. N. (2001). Useful fungi of the world: mu-erh<br />

and silver ears. Mycologist, 15, 19 20.<br />

Pegler, D. N. & Young, T. W. K. (1971). Basidiospore<br />

morphology in the Agaricales. Nova Hedwigia Beihefte,<br />

35, 1 210.<br />

Pegler, D. N. & Young, T. W. K. (1979). The gastroid<br />

Russulales. Transactions of the British Mycological<br />

Society, 72, 353 388.<br />

Pegler, D. N., Spooner, B. M. & Young, T. W. K. (1993).<br />

British Truffles. A Revision of British Hypogeous <strong>Fungi</strong>.<br />

Kew: Royal Botanic Gardens.<br />

Pegler, D. N., Yao, Y.-J. & Li, Y. (1994). The Chinese<br />

caterpillar fungus. Mycologist, 8, 3 5.<br />

Pegler, D. N., Læssøe, T. & Spooner, B. M. (1995). British<br />

Puffballs, Earthstars and Stinkhorns. Kew: Royal Botanic<br />

Gardens.<br />

Pei, M. H., Royle, D. J. & Hunter, T. (1993). Identity and<br />

host alternation of some willow rusts (Melampsora<br />

spp.) in England. Mycological Research, 97, 845 851.<br />

Pell, J. K., Eilenberg, J., Hajek, A. E. & Steinkraus, D. C.<br />

(2001). Biology and pest management potential of<br />

En<strong>to</strong>mophthorales. In <strong>Fungi</strong> as Biocontrol Agents, ed.<br />

T. M. Butt, C. Jackson & N. Magan. Wallingford, UK:<br />

CABI Publishing, pp. 73 152.<br />

Pepin, R. (1980). Le comportement parasitaire de<br />

Sclerotinia tuberosa (Hedw.) Fuckel sur Anemone<br />

nemorosa L. Étude en microscopie pho<strong>to</strong>nique et<br />

électronique à balayage. Mycopathologia, 72, 89 99.


REFERENCES<br />

781<br />

Percudani, R., Trevisi, A., Zambonelli, A. & Ot<strong>to</strong>nello, S.<br />

(1999). Molecular phylogeny of truffles (Pezizales:<br />

Terfeziaceae, Tuberaceae) derived from nuclear<br />

rDNA sequence analysis. Molecular Genetics and<br />

Evolution, 13, 169 180.<br />

Perea, S., López-Ribot, J. L., Kirkpatrick, W. R., et al.<br />

(2001). Prevalence of molecular mechanisms of<br />

resistance <strong>to</strong> azole antifungal agents in Candida<br />

albicans strains displaying high-level fluconazole<br />

resistance isolated from human immunodeficiency<br />

virus-infected patients. Antimicrobial Agents and<br />

Chemotherapy, 45, 2676 2684.<br />

Perez-Moreno, J. & Read, D. J. (2000). Mobilization and<br />

transfer of nutrients from litter <strong>to</strong> tree seedlings via<br />

the vegetative mycelium of ec<strong>to</strong>mycorrhizal plants.<br />

New Phy<strong>to</strong>logist, 145, 301 309.<br />

Perfect, J. R. (2004). Genetic requirements for virulence<br />

in Cryp<strong>to</strong>coccus neoformans. In The Mycota XII: Human<br />

Fungal Pathogens, ed. J. E. Domer & G. S. Kobayashi.<br />

Berlin: Springer-Verlag, pp. 89 112.<br />

Perfect, S. E., Green, J. R. & O’Connell, R. J. (2001).<br />

Surface characteristics of necrotrophic secondary<br />

hyphae produced by the bean anthracnose fungus,<br />

Colle<strong>to</strong>trichum lindemuthianum. European Journal of Plant<br />

Pathology, 107, 813 819.<br />

Perkins, D. D. (1992). Neurospora: the organism behind<br />

the molecular revolution. Genetics, 130, 687 701.<br />

Perkins, D. D. & Turner, B. C. (1988). Neurospora from<br />

natural populations <strong>to</strong>wards the population<br />

biology of a haploid eukaryote. Experimental Mycology,<br />

12, 91 131.<br />

Perkins, D. D., Radford, A. & Sachs, M. S. (2000). The<br />

Neurospora Compendium. Chromosomal Loci. San Diego:<br />

Academic Press.<br />

Perkins, F. O. (1972). The ultrastructure of holdfasts,<br />

‘rhizoids’, and ‘slime tracks’ in thraus<strong>to</strong>chytriaceous<br />

fungi and Labyrinthula spp. Archiv für Mikrobiologie, 84,<br />

95 118.<br />

Persmark, L. & Nordbring-Hertz, B. (1997). Conidial trap<br />

formation of nema<strong>to</strong>de-trapping fungi in soil and<br />

soil extracts. FEMS Microbiology Ecology, 22, 313 324.<br />

Persson, L., Bødker, L. & Larsson-Wikström, M. (1997).<br />

Prevalence and pathogenicity of foot and root rot<br />

pathogens of pea in Southern Scandinavia. Plant<br />

Disease, 81, 171 174.<br />

Petersen, J., Heitz, M. J. & Hagan, I. M. (1998).<br />

Conjugation in S. pombe: identification of a microtubule-organising<br />

centre, a requirement for microtubules<br />

and a role for Mad2. Current Biology, 8,<br />

963 966.<br />

Petersen, R. H. (1973). Aphyllophorales II. The<br />

Clavariaceae and cantharelloid basidiomycetes.<br />

In The <strong>Fungi</strong>: An Advanced Treatise IVB, ed. G. C.<br />

Ainsworth, F. K. Sparrow & A. S. Sussman. New York:<br />

Academic Press, pp. 351 368.<br />

Petersen, R. H. & Cifuentes, J. (1994). Notes on mating<br />

systems of Auriscalpium vulgare and A. villipes.<br />

Mycological Research, 98, 1427 1430.<br />

Petersen, R. H. & Wu, Q. (1992). Auriscalpium vulgare:<br />

mating system and biological species. Mycosystema, 4,<br />

25 31.<br />

Peterson, R. L. & Currah, R. S. (1990). Synthesis of<br />

mycorrhizae between pro<strong>to</strong>corms of Goodyera repens<br />

(Orchidaceae) and Cera<strong>to</strong>basidium cereale. Canadian<br />

Journal of Botany, 68, 1117 1125.<br />

Peterson, R. L., Bonfante, P., Faccio, A. & Uetake, Y.<br />

(1996). The interface between fungal hyphae and<br />

orchid pro<strong>to</strong>corm cells. Canadian Journal of Botany,<br />

74, 1861 1870.<br />

Peterson, R. L., Uetake, Y. & Zelmer, C. (1998). Fungal<br />

symbiosis with orchid pro<strong>to</strong>corms. Symbiosis, 25,<br />

29 55.<br />

Peterson, R. L., Massicotte, H. B. & Melville, L. H.<br />

(2004). Mycorrhizas: Ana<strong>to</strong>my and Cell Biology.<br />

Ottawa: NRC Press; Wallingford, UK: CABI<br />

Publishing.<br />

Peterson, S. W. (2000a). Phylogenetic analysis of<br />

Penicillium species based on ITS and LSU-rDNA<br />

nucleotide sequences. In Integration of Modern<br />

Taxonomic Methods for Penicillium and Aspergillus<br />

Classification, ed. R. A. Samson & J. I. Pitt.<br />

Amsterdam: Harwood Academic Publishers,<br />

pp. 163 178.<br />

Peterson, S. W. (2000b). Phylogenetic relationships in<br />

Aspergillus based on rDNA sequence analysis. In<br />

Integration of Modern Taxonomic Methods for Penicillium<br />

and Aspergillus Classification, ed. R. A. Samson &<br />

J. I. Pitt. Amsterdam: Harwood Academic Publishers,<br />

pp. 323 355.<br />

Petrini, L. E. & Petrini, O. (1985). Xylariaceous fungi as<br />

endophytes. Sydowia, 38, 216 234.<br />

Petrini, O., Petrini, L. E. & Rodriguez, K. F. (1995).<br />

Xylariaceous endophytes: an exercise in biodiversity.<br />

Fi<strong>to</strong>patalogia Brasiliera, 20, 531 539.<br />

Pey<strong>to</strong>n, G. A. & Bowen, C. C. (1963). The<br />

host parasite interface of Peronospora manshurica<br />

on Glycine max. American Journal of Botany, 50,<br />

787 797.<br />

Pfister, D. H. (1997). Cas<strong>to</strong>r, Pollux and life his<strong>to</strong>ries of<br />

fungi. Mycologia, 89, 1 23.<br />

Pfister, D. H. & Kimbrough, J. W. (2001). Discomycetes.<br />

In The Mycota VIIA: Systematics and Evolution, ed.<br />

D. J. McLaughlin, E. J. McLaughlin & P. A. Lemke.<br />

Berlin: Springer-Verlag, pp. 257 281.


782 REFERENCES<br />

Pfunder, M. & Roy, B. A. (2000). Pollina<strong>to</strong>r-mediated<br />

interactions between a pathogenic fungus, Uromyces<br />

pisi (Pucciniaceae), and its host plant, Euphorbia<br />

cyparissias (Euphorbiaceae). American Journal of Botany,<br />

87, 48 55.<br />

Pfunder, M., Schürch, S. & Roy, B. A. (2001). Sequence<br />

variation and geographic distribution of pseudoflower-forming<br />

rust fungi (Uromyces pisi s. lat.)<br />

on Euphorbia cyparissias. Mycological Research, 105,<br />

57 66.<br />

Pfyffer, G. E., Pfyffer, B. U. & Rast, D. M. (1986). The<br />

polyol pattern, chemotaxonomy, and phylogeny of<br />

the fungi. Sydowia, 39, 160 201.<br />

Phaff, H. J. (1990). Isolation of yeasts from natural<br />

sources. In Isolation of Biotechnological Organisms from<br />

Nature, ed. D. P. Labeda. New York: McGraw-Hill<br />

Publishing Co., pp. 53 79.<br />

Phaff, H. J. & Starmer, W. T. (1987). Yeasts associated<br />

with plants, insects and soil. In The Yeasts 1: Biology of<br />

Yeasts, ed. A. H. Rose & J. S. Harrison. London:<br />

Academic Press, pp. 123 180.<br />

Piepenbring, M., Bauer, R. & Oberwinkler, F. (1998a).<br />

Teliospores of smut fungi. General aspects of<br />

teliospore walls and sporogenesis. Pro<strong>to</strong>plasma, 204,<br />

155 169.<br />

Piepenbring, M., Bauer, R. & Oberwinkler, F. (1998b).<br />

Teliospores of smut fungi. Teliospore walls and the<br />

development of ornamentation studied by electron<br />

microscopy. Pro<strong>to</strong>plasma, 204, 170 201.<br />

Piepenbring, M., Hagedorn, G. & Oberwinkler, F.<br />

(1998c). Spore liberation and dispersal in smut fungi.<br />

Botanica Acta, 111, 444 460.<br />

Piérard-Franchimond, C., Hermanns, J. F., Degreef, H. &<br />

Piérard, G. E. (2000). From axioms <strong>to</strong> new insights<br />

in<strong>to</strong> dandruff. Derma<strong>to</strong>logy, 200, 93 98.<br />

Pine, E. M., Hibbett, D. S. & Donoghue, M. J. (1999).<br />

Phylogenetic relationships of cantharelloid and<br />

clavarioid homobasidiomycetes based on mi<strong>to</strong>chondrial<br />

and nuclear rDNA sequences. Mycologia, 91,<br />

944 963.<br />

Pirozynski, K. A. & Dalpé, Y. (1989). Geological his<strong>to</strong>ry of<br />

the Glomaceae with particular reference <strong>to</strong> mycorrhizal<br />

symbiosis. Symbiosis, 7, 1 36.<br />

Pirozynski, K. A. & Malloch, D. W. (1975). The origin of<br />

land plants: a matter of mycotrophism. Biosystems, 6,<br />

153 164.<br />

Pitt, J. I. (1979). The Genus Penicillium and its Teleomorphic<br />

States Eupenicillium and Talaromyces. London:<br />

Academic Press.<br />

Pitt, J. I. (2000). A Labora<strong>to</strong>ry Guide <strong>to</strong> Common Penicillium<br />

Species, 3rd edn. North Ryde, Australia: Food Science<br />

Australia.<br />

Pitt, J. I. (2002). Biology and ecology of <strong>to</strong>xigenic<br />

Penicillium species. Advances in Experimental Medicine<br />

and Biology, 504, 29 41.<br />

Pitt, J. I. & Hocking, A. D. (1985). <strong>Fungi</strong> and Food Spoilage.<br />

Sydney: Academic Press Australia.<br />

Pitt, J. I., Samson, R. A. & Frisvad, J. C. (2000). List of<br />

accepted species and their synonyms in the family<br />

Trichocomaceae. In Integration of Modern Taxonomic<br />

Methods for Penicillium and Aspergillus Classification, ed.<br />

R. A. Samson & J. I. Pitt. Amsterdam: Harwood<br />

Academic Publishers, pp. 9 49.<br />

Plattner, J. J. & Rapoport, H. (1971). The synthesis of d-<br />

and l-sirenin in their absolute configurations. Journal<br />

of the American Chemical Society, 90, 1758 1761.<br />

Pöder, R. (1983). Über Optimierungsstrategien des<br />

Basidiomycetenhymenophors. Morphologisch<br />

phylogenetische Aspekte. Sydowia, 36, 240 251.<br />

Pöder, R. & Kirchmair, M. (1995). Gills and pores; the<br />

impact of geometrical constraints on form, size and<br />

number of basidia. Documents Mycologiques, 100,<br />

337 348.<br />

Pöggeler, S., Risch, S., Kück, U. & Osiewacz, H. D. (1997).<br />

Mating-type genes from the homothallic fungus<br />

Sordaria macrospora are functionally expressed in a<br />

heterothallic ascomycete. Genetics, 147, 567 580.<br />

Poinar, G. O. & Thomas, G. M. (1982). An en<strong>to</strong>mophthoralean<br />

fungus from Dominican amber.<br />

Mycologia, 74, 332 334.<br />

Poinar, G. O. & Thomas, G. M. (1984). A fossil en<strong>to</strong>mogenous<br />

fungus from Dominican amber. Experientia,<br />

40, 578 579.<br />

Polachek, I., Platt, Y. & Aronovitch, J. (1990).<br />

Catecholamines and virulence of Cryp<strong>to</strong>coccus neoformans.<br />

Infection and Immunity, 58, 2919 2922.<br />

Polak, E., Hermann, R., Kües, U. & Aebi, M. (1997).<br />

Asexual sporulation in Coprinus cinereus: structure<br />

and development of oidiophores and oidia in an<br />

Amut Bmut homokaryon. Fungal Genetics and Biology,<br />

22, 112 126.<br />

Põldmaa, K. (2000). Generic delimitation of the<br />

fungicolous Hypocreaceae. Studies in Mycology, 45,<br />

83 94.<br />

Polley, R. W. & Clarkson, J. D. S. (1980). Take-all severity<br />

and yield in winter wheat: relationship established<br />

using a single plant assessment method. Plant<br />

Pathology, 29, 110 116.<br />

Poloczek, E. & Webster, J. (1994). Conidial traps in<br />

Nema<strong>to</strong>c<strong>to</strong>nus (nema<strong>to</strong>phagous basidiomycetes). Nova<br />

Hedwigia, 59, 201 205.<br />

Polonelli, L., Conti, S. & Gerloni, M. (1991). Antibiobodies:<br />

antibiotic-like antiidiotypic antibodies. Journal of<br />

Medical and Veterinary Mycology, 29, 235 242.


REFERENCES<br />

783<br />

Pommer, E.-H. (1995). Morpholine fungicides and<br />

related compounds. In Modern Selective <strong>Fungi</strong>cides.<br />

Properties, Applications, Mechanisms of Action, 2nd edn,<br />

ed. H. Lyr. New York: Gustav Fischer Verlag,<br />

pp. 163 83.<br />

Pommerville, J. & Fuller, M. S. (1976). The cy<strong>to</strong>logy of<br />

the gametes and fertilization of Allomyces macrogynus.<br />

Archiv für Mikrobiologie, 109, 21 30.<br />

Pommerville, J. & Olson, L. W. (1987). Evidence for maleproduced<br />

pheromone in Allomyces macrogynus.<br />

Experimental Mycology, 11, 245 248.<br />

Pontecorvo, G. (1956). The parasexual cycle in fungi.<br />

Annual Review of Microbiology, 10, 393 400.<br />

Poon, H. & Day, A. W. (1974). ‘Fimbriae’ in the fungus<br />

Ustilago violacea. Nature, 250, 648 649.<br />

Poppe, J. (2000). Use of agricultural waste materials in<br />

the cultivation of mushrooms. In Science and<br />

Cultivation of Edible <strong>Fungi</strong>, ed. L. J. L. D. van Griensven.<br />

Rotterdam: A. Balkema, pp. 3 23.<br />

Porter, C. A. & Jaworski, E. G. (1966). The synthesis of<br />

chitin by particulate preparations of Allomyces<br />

macrogynus. Biochemistry, 5, 1149 1154.<br />

Porter, D. (1972). Cell division in the marine slime<br />

mold, Labyrinthula sp., and the role of the bothrosome<br />

in extracellular membrane production.<br />

Pro<strong>to</strong>plasma, 74, 427 448.<br />

Porter, D. (1990). Phylum Labyrinthulomycota. In<br />

Handbook of Pro<strong>to</strong>ctista, ed. L. Margulis, J. O. Corliss, M.<br />

Melkonian & D. J. Chapman. Bos<strong>to</strong>n: Jones & Bartlett,<br />

pp. 388 398.<br />

Porter, D. & Lingle, W. L. (1992). Endolithic thraus<strong>to</strong>chytrid<br />

marine fungi from planted shell fragments.<br />

Mycologia, 84, 289 299.<br />

Portman, R., Moseman, R. & Levetin, E. (1997).<br />

Ultrastructure of basidiospores in North American<br />

members of the genus Calvatia. Mycotaxon, 62,<br />

435 443.<br />

Powell, K. A. & Rayner, A. D. M. (1983). Ultrastructure of<br />

the rhizomorph apex in Armillaria bulbosa in relation<br />

<strong>to</strong> mucilage production. Transactions of the British<br />

Mycological Society, 81, 529 534.<br />

Powell, M. J. (1983). Localization of antimonatemediated<br />

precipitates of cations in zoospores of<br />

Chytriomyces hyalinus. Experimental Mycology, 7,<br />

266 277.<br />

Powell, M. J. (1993). Looking at mycology with a Janus<br />

face: a glimpse at Chytridiomycetes active in the<br />

environment. Mycologia, 85, 1 20.<br />

Powell, M. J. & Roychoudhury, S. (1992). Ultrastructural<br />

organization of Rhizophlyctis harderi zoospores and<br />

redefinition of type 1 microbody lipid globule<br />

complex. Canadian Journal of Botany, 70, 750 761.<br />

Powers, H. & Kuhlman, E. G. (1997). Rusts of hard pines:<br />

fusiform rust. In Compendium of Conifer Diseases, ed.<br />

E. M. Hansen & K. J. Lewis. St. Paul, USA: APS Press,<br />

pp. 27 9.<br />

Pratt, J. E., Johansson, M. & Hüttermann, A. (1998).<br />

Chemical control of Heterobasidion annosum. In<br />

Heterobasidion annosum. Biology, Ecology, Impact and<br />

Control, ed. S. Woodward, J. Stenlid, R. Karjalainan &<br />

A. Hüttermann. Wallingford, UK: CAB International,<br />

pp. 259 82.<br />

Preisig, O., Moleleki, N., Smit, W. A., Wingfield, D. B. &<br />

Wingfield, M. J. (2000). A novel RNA mycovirus in a<br />

hypovirulent isolate of the plant pathogen Diaporthe<br />

ambigua. Journal of General Virology, 81, 3107 3114.<br />

Premdas, P. D. & Kendrick, B. (1991). Colonization of<br />

autumn-shed leaves by four aero-aquatic fungi.<br />

Mycologia, 93, 317 321.<br />

Pring, R. J., Nash, C., Zakaria, M. & Bailey, J. A. (1995).<br />

Infection process and host range of Colle<strong>to</strong>trichum<br />

capsici. Physiological and Molecular Plant Pathology, 46,<br />

137 152.<br />

Pringle, A., Patek, S. N., Fischer, M., S<strong>to</strong>lze, J. & Money,<br />

N. P. (2005). The captured launch of a ballis<strong>to</strong>spore.<br />

Mycologia, 97, 866 871.<br />

Pringle, J. R., Jones, E. W. & Broach, J. R., eds. (1997). The<br />

Molecular and Cellular Biology of the Yeast Saccharomyces,<br />

3. Cold Spring Harbor and New York: Cold Spring<br />

Harbor Labora<strong>to</strong>ry Press.<br />

Prins, T. W., Tudzynski, P., von Tiedemann, A., et al.<br />

(2000). Infection strategies of Botrytis cinerea and<br />

related necrotrophic pathogens. In Fungal Pathology,<br />

ed. J. W. Kronstad. Dordrecht: Kluwer Academic<br />

Publishers, pp. 33 64.<br />

Prusky, D., Freeman, S. & Dickman, M. B., eds. (2000).<br />

Colle<strong>to</strong>trichum. Host Specificity, Pathology, and<br />

Host Pathogen Interaction. St. Paul, USA: APS Press.<br />

Pruyne, D. & Bretscher, A. (2000a). Polarization of cell<br />

growth in yeast. I. Establishment and<br />

maintenance of polarity states. Journal of Cell Science,<br />

113, 365 375.<br />

Pruyne, D. & Bretscher, A. (2000b). Polarization of cell<br />

growth in yeast. II. The role of the cortical actin<br />

cy<strong>to</strong>skele<strong>to</strong>n. Journal of Cell Science, 113, 571 585.<br />

Pruyne, D., Legesse-Miller, A., Gao, L., Dong, Y. &<br />

Bretscher, A. (2004). Mechanisms of polarized growth<br />

and organelle segregation in yeast. Annual Review of<br />

Cell and Developmental Biology, 20, 559 591.<br />

Pryor, B. M. & Gilbertson, R. L. (2000). Molecular<br />

phylogenetic relationships among Alternaria species<br />

and related fungi based upon analysis of nuclear ITS<br />

and mt SSU rDNA sequences. Mycological Research,<br />

104, 1312 1321.


784 REFERENCES<br />

Pukatzki, S., Kessin, R. H. & Mekalanos, J. J. (2002). The<br />

human pathogen Pseudomonas aeruginosa utilizes<br />

conserved virulence pathways <strong>to</strong> infect the social<br />

amoeba Dictyostelium discoideum. Proceedings of the<br />

National Academy of Sciences of the USA, 99, 3159 3164.<br />

Pukkila, P. J. & Cassel<strong>to</strong>n, L. A. (1991). Molecular<br />

genetics of the agaric Coprinus cinereus. In More Gene<br />

Manipulations in <strong>Fungi</strong>, ed. J. W. Bennett & L. L. Lasure.<br />

New York: Academic Press, pp. 126 150.<br />

Punja, Z. K., Chittaranjan, S. & Gaye, M. M. (1992).<br />

Development of black root rot caused by Chalara<br />

elegans on fresh market carrots. Canadian Journal of<br />

Plant Pathology, 14, 299 309.<br />

Purdy, L. H. (1979). Sclerotinia sclerotiorum: his<strong>to</strong>ry,<br />

diseases and symp<strong>to</strong>ma<strong>to</strong>logy, host range,<br />

geographic distribution, and impact. Phy<strong>to</strong>pathology,<br />

69, 875 880.<br />

Purdy, L. H. & Schmidt, R. A. (1996). Status of cacao<br />

witches’ broom: biology, epidemiology and management.<br />

Annual Review of Phy<strong>to</strong>pathology, 34, 573 594.<br />

Purdy, L. H., Kendrick, E. L., Hoffmann, J. A. & Hol<strong>to</strong>n,<br />

C. S. (1963). Dwarf bunt of wheat. Annual Review of<br />

Microbiology, 17, 199 222.<br />

Purdy, L. H., Krupa, S. V. & Dean, J. L. (1985).<br />

<strong>Introduction</strong> of sugarcane rust in<strong>to</strong> the Americas<br />

and its spread <strong>to</strong> Florida. Plant Disease, 69, 689 693.<br />

Purvis, O. W., Coppins, B. J., Hawksworth, D. L., James,<br />

P. W. & Moore, D. M., eds. (1992). The Lichen Flora of<br />

Great Britain and Ireland. London: Natural His<strong>to</strong>ry<br />

Museum.<br />

Raajmakers, J. M. & Weller, D. M. (1998). Natural plant<br />

protection by 2,4-diacetylphloroglucinol-producing<br />

Pseudomonas spp. in take-all decline soils. Molecular<br />

Plant Microbe Interactions, 11, 144 152.<br />

Rabinovich, M. L., Bolobova, A. V. & Vasil’chenko, L. G.<br />

(2004). Fungal decomposition of natural aromatic<br />

structures and xenobiotics: a review. Applied<br />

Biochemistry and Microbiology, 40, 1 17.<br />

Radford, A., S<strong>to</strong>ne, P. J. & Taleb, F. (1996). Cellulase and<br />

amylase complexes. In The Mycota III: Biochemistry and<br />

Molecular Biology, ed. R. Brambl & G. A. Marzluf.<br />

Berlin: Springer-Verlag, pp. 269 294.<br />

Raghavendra Rao, N. N. & Pavgi, M. S. (1993). Life<br />

his<strong>to</strong>ry of Synchytrium species parasitic on<br />

Cucurbitaceae. Indian Phy<strong>to</strong>pathology, 46, 36 43.<br />

Raghukumar, S. (2002). Ecology of the marine protists,<br />

the Labyrinthulomycetes (Thraus<strong>to</strong>chytrids and<br />

Labyrinthulids). European Journal of Protis<strong>to</strong>logy, 38,<br />

127 145.<br />

Raguso, R. A. & Roy, B. A. (1998). ‘Floral’ scent production<br />

by Puccinia rust fungi that mimic flowers.<br />

Molecular Ecology, 7, 1127 1136.<br />

Rainieri, S., Zambonelli, C. & Kaneko, Y. (2003).<br />

Saccharomyces sensu stric<strong>to</strong>: systematics, genetic diversity<br />

and evolution. Journal of Bioscience and<br />

Bioengineering, 96, 1 9.<br />

Raju, N. B. (1992a). Genetic control of the sexual cycle<br />

in Neurospora. Mycological Research, 96, 241 262.<br />

Raju, N. B. (1992b). Functional heterothallism resulting<br />

from homokaryotic conidia and ascospores in<br />

Neurospora tetrasperma. Mycological Research, 96,<br />

103 116.<br />

Raju, N. B. & Lu, B. C. (1970). Meiosis in Coprinus. III.<br />

Timing meiotic events in C. lagopus (sensu Buller).<br />

Canadian Journal of Botany, 48, 2183 2186.<br />

Raju, N. B. & Perkins, D. D. (1994). Diverse programs of<br />

ascus development in pseudohomothallic species of<br />

Neurospora, Gelasinospora and Podospora. Developmental<br />

Genetics, 15, 104 118.<br />

Rambold, G. & Triebel, D. (1992). The Inter-Lecanoralean<br />

Associations. Bibliotheca Lichenologica, 48. Stuttgart:<br />

Gebr. Borntraeger.<br />

Rambourg, A., Jackson, C. L. & Clermont, Y. (2001).<br />

Three dimensional configuration of the secre<strong>to</strong>ry<br />

pathway and segregation of secretion granules in<br />

the yeast Saccharomyces cerevisiae. Journal of Cell Science,<br />

114, 2231 2239.<br />

Ramírez, C. (1982). Manual and Atlas of the Penicillia. New<br />

York: Elsevier Biomedical Press.<br />

Ramos, S., Garcia Acha, I. & Peberdy, J. F. (1975). Wall<br />

structure and the budding process in Pullularia<br />

pullulans. Transactions of the British Mycological Society,<br />

64, 283 288.<br />

Ramsay, L. M. & Gadd, G. M. (1997). Mutants of<br />

Saccharomyces cerevisiae defective in vacuolar function<br />

confirm a role for the vacuole in <strong>to</strong>xic metal ion<br />

de<strong>to</strong>xification. FEMS Microbiology Letters, 152,<br />

293 298.<br />

Ramsbot<strong>to</strong>m, J. (1938). Dry rot in ships. Essex Naturalist,<br />

26, 231 267.<br />

Ramsbot<strong>to</strong>m, J. (1953). Mushrooms and Toads<strong>to</strong>ols. A Study<br />

of the Activities of <strong>Fungi</strong>. London: Collins.<br />

Ramstedt, M. (1999). Rust disease on willows<br />

virulence variation and resistance breeding strategies.<br />

Forest Ecology and Management, 121, 101 111.<br />

Rao, K. M. (1994). Rice Blast Disease. Delhi: Daya<br />

Publishing House.<br />

Raper, C. A. (1985). Strategies for mushroom breeding. In<br />

Developmental Biology of Higher <strong>Fungi</strong>, ed. D. Moore,<br />

L. A. Cassel<strong>to</strong>n, D. A. Wood & J. C. Frankland.<br />

Cambridge: Cambridge University Press, pp. 513 528.<br />

Raper, C. A., Raper, J. R. & Miller, R. E. (1972). Genetic<br />

analysis of the life cycle of Agaricus bisporus.<br />

Mycologia, 64, 1088 1117.


REFERENCES<br />

785<br />

Raper, J. R. (1939). Sexual hormones in Achlya. I.<br />

Indicative evidence of a hormonal coordinating<br />

mechanism. American Journal of Botany, 26, 639 650.<br />

Raper, J. R. (1950). Sexual hormones in Achlya. VII. The<br />

hormonal mechanism in homothallic species.<br />

Botanical Gazette, 112, 1 24.<br />

Raper, J. R. (1957). Hormones and sexuality in lower<br />

plants. Symposia of the Society for Experimental Biology,<br />

11, 143 165.<br />

Raper, J. R. (1959). Sexual versatility and evolutionary<br />

processes in fungi. Mycologia, 51, 107 124.<br />

Raper, J. R. (1966). Genetics of Sexuality in Higher <strong>Fungi</strong>.<br />

New York: Ronald Press Co.<br />

Raper, J. R. & Hoffman, R. M. (1974). Schizophyllum<br />

commune. In Handbook of Genetics 1, ed. R. C. King. New<br />

York: Plenum Press, pp. 597 626.<br />

Raper, K. B. (1984). The Dictyostelids. Prince<strong>to</strong>n: Prince<strong>to</strong>n<br />

University Press.<br />

Raper, K. B. & Fennell, D. I. (1965). The Genus Aspergillus.<br />

Baltimore: Williams & Wilkins Company.<br />

Rasmussen, H. N. (2002). Recent developments in the<br />

study of orchid mycorrhiza. Plant and Soil, 244,<br />

149 163.<br />

Rast, D. & Hollenstein, G. O. (1977). Architecture of the<br />

Agaricus bisporus spore wall. Canadian Journal of<br />

Botany, 55, 2251 2262.<br />

Rast, D. M. & Pfyffer, G. E. (1989). Acyclic polyols and<br />

higher taxa of fungi. Botanical Journal of the Linnean<br />

Society, 99, 39 57.<br />

Ratledge, C. (1997). Microbial lipids. In Biotechnology 7:<br />

Products of Secondary Metabolism, 2nd edn., ed. H.<br />

Kleinkauf & H. vonDöhren. Weinheim, Germany:<br />

VCH, pp. 133 197.<br />

Ratner, D. I. & Kessin, R. H. (2000). Meeting report:<br />

Dictyostelium 2000: a conference on the cell and<br />

developmental biology of a social amoeba, Dundee,<br />

Scotland, July 30 August 4, 2000. Protist, 151,<br />

291 297.<br />

Raudaskoski, M. (1972). Occurrence of microtubules in<br />

the hyphae of Schizophyllum commune during intercellular<br />

nuclear migration. Archiv für Mikrobiologie,<br />

86, 91 100.<br />

Raudaskoski, M. & Vauras, R. (1982). Scanning electron<br />

microscope study of fruit body differentiation in<br />

Schizophyllum commune. Transactions of the British<br />

Mycological Society, 78, 475 481.<br />

Raudaskoski, M. & Viitanen, H. (1982). Effects of<br />

aeration and light on fruit-body induction in<br />

Schizophyllum commune. Transactions of the British<br />

Mycological Society, 78, 89 96.<br />

Rausch, T., Mattusch, P. & Hilgenberg, W. (1981).<br />

Influence of clubroot disease on the growth kinetics<br />

of Chinese cabbage. Phy<strong>to</strong>pathologische Zeitschrift, 102,<br />

28 33.<br />

Raybould, A. F., Gray, A. J. & Clarke, R. T. (1998).<br />

The long-term epidemic of Claviceps purpurea on<br />

Spartina anglica in Poole Harbour: pattern of<br />

infection, effects on seed production and the role<br />

of Fusarium heterosporum. New Phy<strong>to</strong>logist, 138,<br />

495 507.<br />

Rayner, A. D. M. (1991a). The challenge of the individualistic<br />

mycelium. Mycologia, 83, 48 71.<br />

Rayner, A. D. M. (1991b). The phy<strong>to</strong>pathological significance<br />

of mycelial individualism. Annual Review of<br />

Phy<strong>to</strong>pathology, 29, 305 323.<br />

Rayner, A. D. M. & Boddy, L. (1988). Fungal Decomposition<br />

of Wood. Its Biology and Ecology. Chichester, UK: John<br />

Wiley.<br />

Rayner, A. D. M. & Todd, N. K. (1977). Intraspecific<br />

antagonism in natural populations of wood-decaying<br />

basidiomycetes. Journal of General Microbiology,<br />

103, 85 90.<br />

Rayner, A. D. M. & Todd, N. K. (1979). Population and<br />

community structure and dynamics of fungi in<br />

decaying wood. Advances in Botanical Research, 7,<br />

333 420.<br />

Rayner, A. D. M., Powell, K. A., Thompson, W. &<br />

Jennings, D. H. (1985). Morphogenesis of vegetative<br />

organs. In Developmental Biology of Higher <strong>Fungi</strong>, ed.<br />

D. Moore, L. A. Cassel<strong>to</strong>n, D. A. Wood &<br />

J. C. Frankland. Cambridge: Cambridge University<br />

Press, pp. 249 279.<br />

Raziq, F. (2000). Biological and integrated control of<br />

Armillaria root rot. In Armillaria Root Rot: Biology and<br />

Control of Honey Fungus, ed. R. T. V. Fox. Andover, UK:<br />

Intercept Ltd, pp. 183 201.<br />

Read, D. J. (1991). Mycorrhizas in ecosystems nature’s<br />

response <strong>to</strong> the ‘law of the minimum’. In Frontiers in<br />

Mycology, ed. D. L. Hawksworth. Wallingford, UK: CAB<br />

International, pp. 101 130.<br />

Read, D. J. (1996). The structure and function of the<br />

ericoid mycorrhizal root. Annals of Botany N.S., 77,<br />

365 374.<br />

Read, N. D. & Beckett, A. (1985). The ana<strong>to</strong>my of the<br />

mature perithecium in Sordaria humana and its<br />

significance for fungal multicellular development.<br />

Canadian Journal of Botany, 63, 281 296.<br />

Read, N. D. & Beckett, A. (1996). Ascus and<br />

ascospore morphogenesis. Mycological Research, 100,<br />

1281 1314.<br />

Read, S. J., Moss, S. T. & Jones, E. B. G. (1991).<br />

Attachment studies of aquatic hyphomycetes.<br />

Philosophical Transactions of the Royal Society, 334,<br />

449 457.


786 REFERENCES<br />

Read, S. J., Moss, S. T. & Jones, E. B. G. (1992a).<br />

Germination and development of attachment structures<br />

by conidia of aquatic hyphomycetes: light<br />

microscope studies. Canadian Journal of Botany, 70,<br />

831 837.<br />

Read, S. J., Moss, S. T. & Jones, E. B. G. (1992b).<br />

Germination and development of attachment structures<br />

by conidia of aquatic hyphomycetes: a scanning<br />

electron microscopic study. Canadian Journal of<br />

Botany, 70, 838 845.<br />

Reavy, B., Arif, M., Kashiwazaki, S., Webster, K. D. &<br />

Barker, H. (1995). Immunity <strong>to</strong> pota<strong>to</strong> mop-<strong>to</strong>p virus<br />

in Nicotiana benthamiana plants expressing the coat<br />

protein gene is effective against fungal inoculation<br />

of the virus. Molecular Plant Microbe Interactions, 8,<br />

286 291.<br />

Reddy, M. S. & Kramer, C. L. (1975). A taxonomic<br />

zrevision of the Pro<strong>to</strong>mycetales. Mycotaxon, 3, 1 50.<br />

Reddy, P. V., Patel, R. & White, J. F. (1998). Phylogenetic<br />

and developmental evidence supporting reclassification<br />

of cruciferous pathogens Phoma lingam and<br />

Phoma wasabiae in Plenodomus. Canadian Journal of<br />

Botany, 76, 1916 1922.<br />

Redecker, D., Kodner, R. & Graham, L. E. (2000a).<br />

Glomalean fungi from the Ordovician. Science, 289,<br />

1920 1921.<br />

Redecker, D., Mor<strong>to</strong>n, J. B. & Bruns, T. D. (2000b).<br />

Molecular phylogeny of the arbuscular mycorrhizal<br />

fungi Glomus sinuosum and Sclerocystis coremioides.<br />

Mycologia, 92, 282 285.<br />

Redhead, S. A. (1977). The genus Neolecta (Neolectaceae<br />

fam. nov., Lecanorales, Ascomycetes) in Canada.<br />

Canadian Journal of Botany, 55, 301 306.<br />

Redhead, S. A. (2001). Bully for Coprinus a s<strong>to</strong>ry of<br />

manure, minutiae and molecules. McIlvainea, 14,<br />

5 14.<br />

Redhead, S. A., Vilgalys, R., Moncalvo, J.-M., Johnson, J.<br />

& Hopple, J. S. (2001). Coprinus Pers. and the disposition<br />

of Coprinus species sensu la<strong>to</strong>. Taxon, 50, 203 241.<br />

Redlin, S. C. (1991). Discula destructiva sp. nov., cause of<br />

dogwood anthacnose. Mycologia, 83, 633 642.<br />

Redman, R. S., Ranson, J. C. & Rodriguez, R. J. (1999a).<br />

Conversion of the pathogenic fungus Colle<strong>to</strong>trichum<br />

magna <strong>to</strong> a nonpathogenic, endophytic mutualist by<br />

gene disruption. Molecular Plant Microbe Interactions,<br />

12, 969 975.<br />

Redman, R. S., Freeman, S., Clif<strong>to</strong>n, D. R., et al. (1999b).<br />

Biochemical analysis of plant protection afforded by<br />

a nonpathogenic endophytic mutant of<br />

Colle<strong>to</strong>trichum magna. Plant Physiology, 119, 795 804.<br />

Rees, B., Shepherd, V. A. & Ashford, A. E. (1994).<br />

Presence of a motile tubular vacuole system in<br />

different phyla of fungi. Mycological Research, 98,<br />

985 992.<br />

Reeves, F. (1967). The fine structure of ascospore<br />

formation in Pyronema domesticum. Mycologia, 59,<br />

1018 1033.<br />

Reichle, R. E. & Fuller, M. S. (1967). The fine structure of<br />

Blas<strong>to</strong>cladiella emersonii zoospores. American Journal of<br />

Botany, 54, 81 92.<br />

Reichle, R. E. & Lichtwardt, R. W. (1972). Fine structure<br />

of the Trichomycete, Harpella melusinae, from blackfly<br />

guts. Archiv für Mikrobiologie, 81, 103 125.<br />

Reid, D. A. (1974). A monograph of the British<br />

Dacrymycetales. Transactions of the British Mycological<br />

Society, 62, 433 494.<br />

Reijnders, A. F. M. (1963). Problèmes du développement des<br />

carpophores des Agaricales et de quelques groupes voisins.<br />

The Hague: Dr W. Junk.<br />

Reijnders, A. F. M. (1986). Development of the primordium<br />

of the carpophore. In The Agaricales in Modern<br />

Taxonomy, 4th edn, ed. R. Singer. Königstein,<br />

Germany: Koeltz Scientific Books, pp. 20 29.<br />

Reijnders, A. F. M. (2000). A morphogenetic analysis of<br />

the basic characters of the gasteromycetes and their<br />

relation <strong>to</strong> other basidiomycetes. Mycological Research,<br />

104, 900 910.<br />

Reijnders, A. F. M. & Stalpers, J. A. (1992). The development<br />

of the hymenophoral trama in the<br />

Aphyllophorales and the Agaricales. Studies in<br />

Mycology, 34, 1 109.<br />

Reinhardt, M. O. (1892). Das Wachstum der Pilzhyphen.<br />

Ein Beitrag zur Kenntnis des Flächenwachstums<br />

vegetalischer Zellmembranen. Jahrbücher für<br />

Wissenschaftliche Botanik, 23, 479 566.<br />

Reischer, H. S. (1951). Growth of Saprolegniaceae in<br />

synthetic media. I. Inorganic nutrition. Mycologia, 43,<br />

142 155.<br />

Renzel, S., Esselborn, S., Sauer, H. W. & Hildebrandt, A.<br />

(2000). Calcium and malate are sporulation-promoting<br />

fac<strong>to</strong>rs of Physarum polycephalum. Journal of<br />

Bacteriology, 182, 6900 6905.<br />

Reshetnikov, S. V., Wasser, S. P., Nevo, E., Duckman, I. &<br />

Tsukor, K. (2000). Medicinal value of the genus<br />

Tremella Pers. (Heterobasidiomycetes). International<br />

Journal of Medicinal Mushrooms, 2, 169 193.<br />

Restrepo, A., McEwen, J. G. & Castañeda, E. (2001). The<br />

habitat of Paracoccidioides brasiliensis: how far from<br />

solving the riddle? Medical Mycology, 39, 233 241.<br />

Reynaga-Peña, C. G., Gierz, G. & Bartnicki-Garcia, S.<br />

(1997). Analysis of the role of the Spitzenkörper in<br />

fungal morphogenesis by computer simulation of<br />

apical branching in Aspergillus niger. Proceedings of the<br />

National Academy of Sciences of the USA, 94, 9096 9101.


REFERENCES<br />

787<br />

Reynolds, D. R. (1971). Wall structure of a bitunicate<br />

ascus. Planta, 98, 244 257.<br />

Reynolds, D. R. (1989). The bitunicate ascus paradigm.<br />

Botanical Review, 55, 1 52.<br />

Reynolds, D. R. & Taylor, J. W., eds. (1993). The Fungal<br />

Holomorph: Mi<strong>to</strong>tic, Meiotic and Pleomorphic Speciation in<br />

Fungal Systematics. Wallingford, UK: CAB<br />

International.<br />

Rghei, N. A., Castle, A. J. & Manocha, M. S. (1992).<br />

Involvement of fimbriae in fungal host parasite<br />

interaction. Physiological and Molecular Plant Pathology,<br />

41, 139 148.<br />

Ribeiro, O. K. (1983). Physiology of asexual<br />

sporulation and spore germination in Phy<strong>to</strong>phthora.<br />

In Phy<strong>to</strong>phthora. Its Biology, Taxonomy, Ecology,<br />

and Pathology, ed. D. C. Erwin,<br />

S. Bartnicki-Garcia & P. H. Tsao. St. Paul: APS Press,<br />

pp. 55 70.<br />

Ribes, J. A., Vanover-Sams, C. L. & Baker, D. J. (2000).<br />

Zygomycetes in human disease. Clinical Microbiology<br />

Reviews, 13, 236 301.<br />

Rice, A. V. & Currah, R. S. (2005). Oidiodendron: a<br />

survey of the named species and related<br />

anamorphs of Myxotrichum. Studies in Mycology,<br />

53, 83 120.<br />

Rice, S. L., Beuchat, L. R. & Worthing<strong>to</strong>n, R. E. (1977).<br />

Patulin production by Byssochlamys spp. in fruit<br />

juices. Applied and Environmental Microbiology, 34,<br />

791 796.<br />

Richards, W. C. (1993). Cera<strong>to</strong>-ulmin: a unique wilt<br />

disease <strong>to</strong>xin of instrumental significance in the<br />

development of Dutch Elm disease. In Dutch Elm<br />

Disease Research: Cellular and Molecular Approaches, ed.<br />

M. B. Sticklen & J. L. Sherald. Berlin: Springer-Verlag,<br />

pp. 89 151.<br />

Richardson, D. H. S. (1975). The Vanishing Lichens. Their<br />

His<strong>to</strong>ry, Biology and Importance. New<strong>to</strong>n Abbot: David<br />

& Charles.<br />

Richardson, D. H. S. (1988). Medicinal and other<br />

economic aspects of lichens. In CRC Handbook of<br />

Lichenology III, ed. M. Galun. Boca Ra<strong>to</strong>n: CRC Press,<br />

pp. 93 108.<br />

Richardson, D. H. S. (1991). Lichens and man. In Frontiers<br />

in Mycology, ed. D. L. Hawksworth. Wallingford, UK:<br />

CAB International, pp. 187 210.<br />

Richardson, D. H. S. (1999). War in the world of lichens:<br />

parasitism and symbiosis as exemplified by lichens<br />

and lichenicolous fungi. Mycological Research, 103,<br />

641 650.<br />

Richardson, M. J. (1972). Coprophilous ascomycetes on<br />

different dung types. Transactions of the British<br />

Mycological Society, 58, 37 48.<br />

Richardson, M. J. (2001). Diversity and occurrence of<br />

coprophilous fungi. Mycological Research, 105,<br />

387 402.<br />

Richardson, M. J. & Leadbeater, G. (1972). Pip<strong>to</strong>cephalis<br />

fimbriata sp. nov., and observations on the occurrence<br />

of Pip<strong>to</strong>cephalis and Syncephalis. Transactions of<br />

the British Mycological Society, 58, 205 215.<br />

Richardson, M. J. & Watling, R. (1997). Keys <strong>to</strong> <strong>Fungi</strong> on<br />

Dung. S<strong>to</strong>urbridge, UK: British Mycological Society.<br />

Richter, S. G. & Barnard, J. (2002). The radiation<br />

resistance of ascospores and sclerotia of Pyronema<br />

domesticum. Journal of Industrial Microbiology and<br />

Biotechnology, 29, 51 54.<br />

Riehl, R. M., Toft, D. O., Meyer, M. D., Carlson, G. L. &<br />

McMorris, T. C. (1984). Detection of a pheromonebinding<br />

protein in the aquatic fungus Achlya<br />

ambisexualis. Experimental Cell Research, 153, 544 549.<br />

Riethmüller, A., Weiss, M. & Oberwinkler, F. (1999).<br />

Phylogenetic studies of Saprolegniomycetidae and<br />

related groups based on nuclear large subunit<br />

ribosomal DNA sequences. Canadian Journal of Botany,<br />

77, 1790 1800.<br />

Riethmüller, A., Voglmayr, H., Göker, M., Weiss, M. &<br />

Oberwinkler, F. (2002). Phylogenetic relationships of<br />

the downy mildews (Peronosporales) and related<br />

groups based on nuclear large subunit ribosomal<br />

DNA sequences. Mycologia, 94, 834 849.<br />

Rifai, M. A. & Webster, J. (1966). Culture studies on<br />

Hypocrea and Trichoderma. III. H. lactea (¼ H. citrina)<br />

and H. pulvinata. Transactions of the British Mycological<br />

Society, 49, 297 310.<br />

Rikkinen, J. (2003). Ecological and evolutionary role of<br />

pho<strong>to</strong>biont-mediated guilds in lichens. Symbiosis, 34,<br />

99 110.<br />

Rikkinen, J. & Poinar, G. (2000). A new species of<br />

resinicolous Chaenothecopsis (Mycocaliciaceae,<br />

Ascomycota) from 20 million year old Bitterfeld<br />

amber, with remarks on the biology of resinicolous<br />

fungi. Mycological Research, 104, 7 15.<br />

Rikkinen, J., Oksanen, I. & Lohtander, K. (2002). Lichen<br />

guilds share related cyanobacterial symbionts.<br />

Science, 297, 357.<br />

Rinaldi, M. G. (1989). Zygomycosis. Infectious Disease<br />

Clinics of North America, 3, 19 41.<br />

Riousset, L., Riousset, G., Chevalier, G. & Bardet, M. C.<br />

(2001). Truffes d’Europe et de la Chine. Paris: INRA.<br />

Riquelme, M., Reynaga-Peña, C. G., Gierz, G. &<br />

Bartnicki-García, S. (1998). What determines growth<br />

direction in fungal hyphae? Fungal Genetics and<br />

Biology, 24, 101 109.<br />

Rishbeth, J. (1952). Control of Fomes annosus Fr. Forestry,<br />

25, 41 50.


788 REFERENCES<br />

Rishbeth, J. (1961). Inoculation of pine stumps against<br />

Fomes annosus. Nature, 191, 826 827.<br />

Rishbeth, J. (1968). The growth rate of Armillaria mellea.<br />

Transactions of the British Mycological Society, 51,<br />

575 586.<br />

Ristaino, J. B. (2002). Tracking his<strong>to</strong>ric migrations of<br />

the Irish pota<strong>to</strong> famine pathogen, Phy<strong>to</strong>phthora<br />

infestans. Microbes and Infection, 4, 1369 1377.<br />

Rivas, S. & Thomas, C. M. (2005). Molecular interactions<br />

between <strong>to</strong>ma<strong>to</strong> and the leaf mold pathogen<br />

Cladosporium fulvum. Annual Review of Phy<strong>to</strong>pathology,<br />

43, 395 436.<br />

Robb, J. & Lee, B. (1986a). Developmental sequence of<br />

the attack apparatus of Hap<strong>to</strong>glossa mirabilis.<br />

Pro<strong>to</strong>plasma, 135, 102 111.<br />

Robb, J. & Lee, B. (1986b). Ultrastructure of mature and<br />

fired gun cells of Hap<strong>to</strong>glossa mirabilis. Canadian<br />

Journal of Botany, 64, 1935 1947.<br />

Roberts, D. R., Mims, C. W. & Fuller, M. S. (1996).<br />

Ultrastructure of the ungerminated conidium of<br />

Blumeria graminis f. sp. hordei. Canadian Journal of<br />

Botany, 74, 231 237.<br />

Roberts, P. (1999). Rhizoc<strong>to</strong>nia-Forming <strong>Fungi</strong>. A Taxonomic<br />

Guide. Kew: Royal Botanic Gardens.<br />

Roberts, R. G., Reymond, S. T. & Andersen, B. (2000).<br />

RAPD fragment pattern analysis and morphological<br />

segregation of small-spored Alternaria species<br />

and species groups. Mycological Research, 104,<br />

151 160.<br />

Robertson, N. F. (1965). Presidential address. The fungal<br />

hypha. Transactions of the British Mycological Society, 48,<br />

1 8.<br />

Robertson, S. K., Bond, D. J. & Read, N. D. (1998).<br />

Homothallism and heterothallism in Sordaria brevicollis.<br />

Mycological Research, 102, 1215 1223.<br />

Robinow, C. F. (1963). Observations on cell growth,<br />

mi<strong>to</strong>sis and division in the fungus Basidiobolus<br />

ranarum. Journal of Cell Biology, 17, 123 152.<br />

Roca, M. G., Davide, L. C. & Mendes-Costa, M. C. (2003).<br />

Cy<strong>to</strong>genetics of Colle<strong>to</strong>trichum lindemuthianum<br />

(Glomerella cingulata f. sp. phaseoli). Fi<strong>to</strong>pa<strong>to</strong>logia<br />

Brasileira, 28, 367 373.<br />

Rochon, D., Kakani, K., Robbins, M. & Reade, R. (2004).<br />

Molecular aspects of plant virus transmission by<br />

Olpidium and plasmodiophorid vec<strong>to</strong>rs. Annual Review<br />

of Phy<strong>to</strong>pathology, 42, 211 241.<br />

Rodriguez, R. J. & Redman, R. S. (1997). Fungal life-styles<br />

and ecosystem dynamics: biological aspects of plant<br />

pathogens, plant endophytes and saprophytes.<br />

Advances in Botanical Research, 24, 169 193.<br />

Rodríguez-Peña, J. M., Cid, V. J., Arroyo, J. & Nombela, C.<br />

(2000). A novel family of cell wall related proteins<br />

regulated differently during the yeast life cycle.<br />

Molecular and Cellular Biology, 20, 3245 3255.<br />

Roelfs, A. P. (1984). Race specificity and methods of<br />

study. In The Cereal Rusts 1: Origins, Specificity, Structure,<br />

and Physiology, ed. W. R. Bushnell & A. P. Roelfs.<br />

Orlando: Academic Press, pp. 131 64.<br />

Roelfs, A. P. (1985). Wheat and rye stem rust. In The<br />

Cereal Rusts 2: Diseases, Distribution, Epidemiology, and<br />

Control, ed. A. P. Roelfs & W. R. Bushnell. Orlando:<br />

Academic Press, pp. 3 37.<br />

Roelfs, A. P. (1986). Development and impact of<br />

regional cereal rust epidemics. In Plant Disease<br />

Epidemiology 1: Population Dynamics and Management,<br />

ed. K. J. Leonard & W. E. Fry. New York: McMillan,<br />

pp. 129 50.<br />

Roelfs, A. P. & Bushnell, W. R., eds. (1985). The Cereal<br />

Rusts 2: Diseases, Distribution, Epidemiology, and Control.<br />

Orlando: Academic Press.<br />

Roelfs, A. P. & Martens, J. W. (1988). An international<br />

system of nomenclature for Puccinia graminis f. sp.<br />

tritici. Phy<strong>to</strong>pathology, 78, 526 533.<br />

Roger, A. J., Smith, M. W., Doolittle, R. F. & Doolittle,<br />

W. F. (1996). Evidence for the Heterolobosea from<br />

phylogenetic analysis of genes encoding glyceraldehyde-3-phosphate<br />

dehydrogenase. Journal of<br />

Eukaryotic Microbiology, 43, 475 485.<br />

Rogers, H. J., Buck, K. W. & Brasier, C. M. (1986).<br />

Transmission of double-stranded RNA and a disease<br />

fac<strong>to</strong>r in Ophios<strong>to</strong>ma ulmi. Plant Pathology, 35, 277 287.<br />

Rogers, J. D. (1965). Hypoxylon fuscum. I. Cy<strong>to</strong>logy of the<br />

ascus. Mycologia, 57, 789 803.<br />

Rogers, J. D. (1975a). Xylaria polymorpha. II. Cy<strong>to</strong>logy of a<br />

form with typical robust stromata. Canadian Journal<br />

of Botany, 53, 1736 1743.<br />

Rogers, J. D. (1975b). Hypoxylon serpens: cy<strong>to</strong>logy and<br />

taxonomic considerations. Canadian Journal of Botany,<br />

53, 52 55.<br />

Rogers, J. D. (1979). The Xylariaceae: systematic, biological<br />

and evolutionary aspects. Mycologia, 71, 1 42.<br />

Rogers, J. D. (2000). Thoughts and musings on tropical<br />

Xylariaceae. Mycological Research, 104, 1412 1420.<br />

Rogerson, C. T. (1970). The Hypocrealean fungi<br />

(Ascomycetes, Hypocreales). Mycologia, 62, 865 910.<br />

Rolinson, G. N. (1998). Forty years of b-lactam research.<br />

Journal of Antimicrobial Chemotherapy, 41, 589 603.<br />

Rollins, J. A. (2003). The Sclerotinia sclerotiorum pac1 gene<br />

is required for sclerotial development and virulence.<br />

Molecular Plant Microbe Interactions, 16, 785 795.<br />

Rollins, J. A. & Dickman, M. B. (2001). pH signalling in<br />

Sclerotinia sclerotiorum: identification of a pacC/RIM1<br />

homolog. Applied and Environmental Microbiology, 67,<br />

75 81.


REFERENCES<br />

789<br />

Roncal, T., Cordobés, S., Sterner, O. & Ugalde, U. (2002).<br />

Conidiation in Penicillium cyclopium is induced by<br />

conidiogenone, an endogenous diterpene. Eukaryotic<br />

Cell, 1, 823 829.<br />

Roper, J. A. (1966). Mechanisms of inheritance. 3. The<br />

parasexual cycle. In The <strong>Fungi</strong>. An Advanced Treatise II:<br />

The Fungal Organism, ed. G. C. Ainsworth & A. S.<br />

Sussman. New York: Academic Press, pp. 589 617.<br />

Rose, C. I. & Hawksworth, D. L. (1981). Lichen recolonization<br />

in London’s cleaner air. Nature, 289, 289 292.<br />

Rosin, I. V. & Moore, D. (1985). Differentiation of the<br />

hymenium in Coprinus cinereus. Transactions of the<br />

British Mycological Society, 84, 621 628.<br />

Rosinski, M. A. (1961). Development of the ascus in<br />

Cera<strong>to</strong>cystis ulmi. American Journal of Botany, 48,<br />

285 293.<br />

Ross, I. K. (1976). Nuclear migration rates in Coprinus<br />

congregatus: a new record? Mycologia, 68, 418 422.<br />

Rossignol, M. & Silar, P. (1996). Genes that control<br />

longevity in Podospora anserina. Mechanisms of Ageing<br />

and Development, 90, 183 193.<br />

Rossman, A. Y. (1983). The phragmosporous species of<br />

Nectria and related genera (Calonectria, Ophionectria,<br />

Paranectria, Scoleconectria and Trichonectria). Mycological<br />

Papers, 150, 1 164.<br />

Rossman, A. Y. (1996). Morphological and molecular<br />

perspectives on systematics of the Hypocreales.<br />

Mycologia, 88, 1 19.<br />

Rossman, A. Y. (2000). Towards monophyletic genera in<br />

the holomorphic Hypocreales. Studies in Mycology, 45,<br />

27 34.<br />

Rossman, A. Y., Samuels, G. J., Rogerson, C. T. & Lowen,<br />

R. (1999). Genera of Bionectriaceae, Hypocreaceae<br />

and Nectriaceae (Hypocreales, Ascomycetes). Studies<br />

in Mycology, 42, 1 248.<br />

Rotem, J. (1994). The Genus Alternaria. Biology,<br />

Epidemiology and Pathogenicity. St Paul, USA: APS Press.<br />

Rothman, J. E. & Orci, L. (1992). Molecular dissection of<br />

the secre<strong>to</strong>ry pathway. Nature, 355, 409 415.<br />

Rouxel, T. & Balesdent, M. H. (2005). The stem canker<br />

(blackleg) fungus, Lep<strong>to</strong>sphaeria maculans, enters<br />

the genomic era. Molecular Plant Pathology, 6,<br />

225 241.<br />

Roy, B. A. (1994). The effects of pathogen-induced<br />

pseudoflowers and buttercups on each other’s insect<br />

visitation. Ecology, 75, 352 358.<br />

Roy, B. & Raguso, R. A. (1997). Olfac<strong>to</strong>ry versus visual<br />

cues in a floral mimicry system. Oecologia, 109,<br />

414 426.<br />

Roy, B. A., Vogler, D. R., Bruns, T. D. & Szaro, T. M. (1998).<br />

Cryptic species in the Puccinia monoica complex.<br />

Mycologia, 90, 846 853.<br />

Rubiales, D. & Niks, R. E. (1996). Avoidance of rust<br />

infection by some genotypes of Hordeum chilense due<br />

<strong>to</strong> their relative inability <strong>to</strong> induce the formation of<br />

appressoria. Physiological and Molecular Plant Pathology,<br />

49, 89 101.<br />

Rubner, A. (1996). Revision of the predacious hyphomycetes<br />

in the Dactylella Monacrosporium complex.<br />

Studies in Mycology, 39, 1 134.<br />

Ruch, D. G. & Motta, J. J. (1987). Ultrastructure and<br />

cy<strong>to</strong>chemistry of dormant basidiospores of Psilocybe<br />

cubensis. Mycologia, 79, 387 398.<br />

Ruch, D. G. & Nurtjahja, K. (1996). The fine structure<br />

and selected his<strong>to</strong>chemistry of ungerminated basidiospores<br />

of Agrocybe acericola. Canadian Journal of<br />

Botany, 74, 780 787.<br />

Rugner, A., Rumbolz, J., Huber, B., et al. (2002).<br />

Formation of overwintering structures of<br />

Uncinula neca<strong>to</strong>r and colonization of<br />

grapevine under field conditions. Plant Pathology,<br />

51, 322 330.<br />

Ruiz-Herrera, J. (1992). Fungal Cell Wall: Structure,<br />

Synthesis and Assembly. Boca Ra<strong>to</strong>n: CRC Press.<br />

Ruiz-Herrera, J., León-Ramírez, C., Cabrera-Ponce,<br />

J. L., Martínez-Espinoza, A. D. & Herrera-Estrella, L.<br />

(1999). Completion of the sexual cycle and<br />

demonstration of genetic recombination in Ustilago<br />

maydis in vitro. Molecular and General Genetics, 262,<br />

468 472.<br />

Rundel, P. W. (1969). Clinal variation in the production<br />

of usnic acid in Cladonia subtenuis along light<br />

gradients. Bryologist, 72, 40 44.<br />

Rupeš, I., Mao, W.-Z., Åström, H. & Raudaskoski, M.<br />

(1995). Effects of nocodazole and brefeldin A on<br />

microtubule cy<strong>to</strong>skele<strong>to</strong>n and membrane organization<br />

in the homobasidiomycete Schizophyllum<br />

commune. Pro<strong>to</strong>plasma, 185, 212 221.<br />

Rush, C. M., Stein, J. M., Bowden, R. L., et al. (2005).<br />

Status of Karnal bunt of wheat in the United States<br />

1996 <strong>to</strong> 2004. Plant Disease, 89, 212 223.<br />

Russell, I. & Stewart, G. G. (1995). Brewing. In<br />

Biotechnology 9: Enzymes, Biomass, Food and Feed, ed.<br />

G. Reed & T. W. Nagodawithana. Weinheim,<br />

Germany: VCH, pp. 419 62.<br />

Ryley, R., Bhuiyan, S., Herde, D. & Gordan, B. (2003).<br />

Efficacy, timing and method of application of<br />

fungicides for management of sorghum ergot<br />

caused by Claviceps africana. Australasian Journal of<br />

Plant Pathology, 32, 329 338.<br />

Saenz, G. S. & Taylor, J. W. (1999a). Phylogeny of the<br />

Erysiphales (powdery mildews) inferred from internal<br />

transcribed spacer ribosomal DNA sequences.<br />

Canadian Journal of Botany, 77, 150 168.


790 REFERENCES<br />

Saenz, G. S. & Taylor, J. W. (1999b). Phylogenetic<br />

relationships of Meliola and Meliolina inferred from<br />

small subunit rRNA sequences. Mycological Research,<br />

103, 1049 1056.<br />

Saikawa, M. & Morikawa, C. (1985). Electron microscopy<br />

on a nema<strong>to</strong>de-trapping fungus,<br />

Acaulopage pec<strong>to</strong>spora. Canadian Journal of Botany,<br />

63, 1386 1390.<br />

Sakagami, Y., Yoshida, M., Isogai, A. & Suzuki, A. (1981).<br />

Peptidal sex hormones inducing conjugation tube<br />

formation in compatible mating-type cells of<br />

Tremella mesenterica. Science, 212, 1525 1527.<br />

Sakai, S., Ka<strong>to</strong>, M. & Nagamasu, H. (2000). Ar<strong>to</strong>carpus<br />

(Moraceae)-gall midge pollination mutualism<br />

mediated by a male-flower parasitic fungus. American<br />

Journal of Botany, 87, 440 445.<br />

Sakaki, H., Nochide, H., Komemushi, S. & Miki, W.<br />

(2002). Effect of active oxygen species on the<br />

productivity of <strong>to</strong>rularhodin by Rhodo<strong>to</strong>rula glutinis<br />

No. 21. Journal of Bioscience and Bioengineering, 93,<br />

338 340.<br />

Salaman, R. N. (1949). The His<strong>to</strong>ry and Social Influence of<br />

the Pota<strong>to</strong>. Cambridge: Cambridge University Press.<br />

Salamati, S., Zhan, J., Burdon, J. J. & McDonald, B. A.<br />

(2000). The genetic structure of field populations of<br />

Rhynchosporium secalis from three continents suggests<br />

moderate gene flow and regular recombination.<br />

Phy<strong>to</strong>pathology, 90, 901 908.<br />

Salas, S. D., Bennett, J. E., Kwon-Chung, K. J., Perfect, J. R.<br />

& Williamson, P. R. (1996). Effect of the laccase gene,<br />

CNLAC1, on virulence of Cryp<strong>to</strong>coccus neoformans.<br />

Journal of Experimental Medicine, 184, 377 386.<br />

Salmeron, J., Vernooij, B., Law<strong>to</strong>n, K., et al. (2002).<br />

Powdery mildew control through transgenic expression<br />

of antifungal proteins, resistance genes, and<br />

systemic acquired resistance. In The Powdery Mildews.<br />

A Comprehensive Treatise, ed. R. R. Bélanger, W. R.<br />

Bushnell, A. J. Dik & T. L. W. Carver. St. Paul, USA: APS<br />

Press, pp. 268 87.<br />

Salvin, S. B. (1941). Comparative studies on the primary<br />

and secondary zoospores of the Saprolegniaceae. I.<br />

Influence of temperature. Mycologia, 53, 592 600.<br />

Samborski, D. J. (1985). Wheat leaf rust. In The Cereal<br />

Rusts 2: Diseases, Distribution, Epidemiology, and Control,<br />

ed. A. P. Roelfs & W. R. Bushnell. Orlando: Academic<br />

Press, pp. 39 59.<br />

Samson, R. A. (2000). List of names of Trichocomaceae<br />

published between 1992 and 1999. In Integration of<br />

Modern Taxonomic Methods for Penicillium and<br />

Aspergillus Classification, ed. R. A. Samson & J. I. Pitt.<br />

Amsterdam: Harwood Academic Publishers,<br />

pp. 73 9.<br />

Samson, R. A., Evans, H. G. & Latgé, J.-P. (1988). Atlas of<br />

En<strong>to</strong>mopathogenic <strong>Fungi</strong>. Berlin: Springer-Verlag.<br />

Samson, R. A., Hoekstra, E. S., Frisvad, J. C. & Filtenborg,<br />

O. (2002). <strong>Introduction</strong> <strong>to</strong> Food-and Airborne <strong>Fungi</strong>, 6th<br />

edn. Utrecht: Centraalbureau voor<br />

Schimmelcultures.<br />

Samson, R. A., Seifert, K. A., Kuijpers, A. F. A.,<br />

Houbraken, J. A. M. P. & Frisvad, J. C. (2004).<br />

Phylogenetic analysis of Penicillium subgenus<br />

Penicillium using partial b-tubulin sequences. Studies<br />

in Mycology, 49, 175 200.<br />

Samuels, G. J. (1983). Ascomycetes of New Zealand. 6.<br />

Atricordyceps harposporifera gen. et sp. nov. and its<br />

Harposporium anamorph. New Zealand Journal of<br />

Botany, 21, 171 176.<br />

Samuels, G. J. (1996). Trichoderma: a review of biology<br />

and systematics of the genus. Mycological Research,<br />

100, 923 935.<br />

Samuels, G. J. (2006). Trichoderma: systematics, the<br />

sexual state, and ecology. Phy<strong>to</strong>pathology, 96,<br />

195 206.<br />

Samuels, G. J. & Blackwell, M. (2001). Pyrenomycetes<br />

fungi with perithecia. In The Mycota VIIA: Systematics<br />

and Evolution, ed. D. J. McLaughlin,<br />

E. G. McLaughlin & P. A. Lemke. Berlin: Springer-<br />

Verlag, pp. 221 55.<br />

Samuels, G. J. & Brayford, D. (1990). Variation in Nectria<br />

radicicola and its anamorph, Cylindrocarpon destructans.<br />

Mycological Research, 94, 433 442.<br />

Samuels, G. J. & Lodge, D. J. (1996). Three species of<br />

Hypocrea with stipitate stromata and Trichoderma<br />

anamorphs. Mycologia, 88, 302 315.<br />

Samuels, G. J., Pardo-Schultheiss, R., Hebbar, K. P., et al.<br />

(2000). Trichoderma stromaticum sp. nov., a parasite of<br />

the cacao witches’ broom. Mycological Research, 104,<br />

760 764.<br />

Samuels, G. J., Nirenberg, H. I. & Seifert, K. A. (2001).<br />

Perithecial species of Fusarium. In Fusarium. Paul E.<br />

Nelson Memorial Symposium, ed. B. O. A. Summerell, J. F.<br />

Leslie, D. Backhouse, W. L. Bryden & L. W. Burgess.<br />

St. Paul, USA: APS Press, pp. 1 14.<br />

Samuels, G. J., Dodd, S. L., Gams, W., Castlebury, L. A. &<br />

Petrini, O. (2002). Trichoderma species associated with<br />

the green mold epidemic of commercially grown<br />

Agaricus bisporus. Mycologia, 94, 146 170.<br />

Samuelson, D. A. (1978a). Asci of the Pezizales. I. The<br />

apical apparatus of iodine-positive species. Canadian<br />

Journal of Botany, 56, 1860 1875.<br />

Samuelson, D. A. (1978b). Asci of the Pezizales. VI. The<br />

apical apparatus of Morchella esculenta, Helvella crispa,<br />

and Rhizina undulata. Canadian Journal of Botany, 56,<br />

3069 3082.


REFERENCES<br />

791<br />

Sánchez, C. (2004). Modern aspects of mushroom<br />

culture technology. Applied Microbiology and<br />

Biotechnology, 64, 756 762.<br />

Sanders, D. (1988). <strong>Fungi</strong>. In Solute Transport in Plant Cells<br />

and Tissues, ed. D. A. Baker & J. L. Hall. Harlow:<br />

Longman, pp. 106 65.<br />

Sanders, P. F. & Webster, J. (1978). Survival of aquatic<br />

hyphomycetes in terrestrial situations. Transactions of<br />

the British Mycological Society, 71, 231 237.<br />

Sanders, P. F. & Webster, J. (1980). Sporulation<br />

responses of some aquatic hyphomycetes in flowing<br />

water. Transactions of the British Mycological Society, 74,<br />

601 605.<br />

Sandmann, G. & Misawa, N. (2002). Fungal carotenoids.<br />

In The Mycota X: Industrial Applications, ed. H. D.<br />

Osiewacz. Berlin: Springer-Verlag, pp. 247 262.<br />

Sangar, V. K. & Dugan, P. R. (1973). Chemical composition<br />

of the cell wall of Smittium culisetae<br />

(Trichomycetes). Mycologia, 65, 421 431.<br />

Sanglard, D. (2002). Clinical relevance of mechanisms<br />

of antifungal drug resistance in yeasts. Enfermedades<br />

Infecciosas Microbiologia Clinica, 20, 462 470.<br />

Sanglard, D. & Bille, J. (2002). Current understanding of<br />

the modes of action of and resistance <strong>to</strong> conventional<br />

and emerging antifungal agents for treatment<br />

of Candida infections. In Candida and Candidiasis, ed.<br />

R. A. Calderone. Washing<strong>to</strong>n: ASM Press,<br />

pp. 349 383.<br />

Sanogo, S., Pomella, A., Hebbar, R. K., et al. (2002).<br />

Production and germination of conidia of<br />

Trichoderma stromaticum, a mycoparasite of Crinipellis<br />

perniciosa on cacao. Phy<strong>to</strong>pathology, 92, 1032 1037.<br />

Sansome, E. (1963). Meiosis in Pythium debaryanum<br />

Hesse and its significance in the life his<strong>to</strong>ry of the<br />

Biflagellatae. Transactions of the British Mycological<br />

Society, 46, 63 72.<br />

Sansome, E. & Sansome, F. W. (1974). Cy<strong>to</strong>logy and lifehis<strong>to</strong>ry<br />

of Peronospora parasitica on Capsella bursapas<strong>to</strong>ris<br />

and of Albugo candida on C. bursa-pas<strong>to</strong>ris and<br />

on Lunaria annua. Transactions of the British Mycological<br />

Society, 62, 323 332.<br />

Saupe, S. J. (2000). Molecular genetics of heterokaryon<br />

incompatibility in filamen<strong>to</strong>us ascomycetes.<br />

Microbiology and Molecular Biology Reviews, 64,<br />

489 502.<br />

Sauter, H., Steglich, W. & Anke, T. (1999). Strobilurins:<br />

evolution of a new class of active substances.<br />

Angewandte Chemie International <strong>Edition</strong>, 38,<br />

1328 1349.<br />

Savile, D. B. O. (1968). Possible interrelationships<br />

between fungal groups. In The <strong>Fungi</strong>, an Advanced<br />

Treatise III: The Fungal Population, ed. G. C. Ainsworth<br />

& A. S. Sussman. New York: Academic Press,<br />

pp. 649 675.<br />

Savile, D. B. O. & Hayhoe, H. N. (1978). The potential<br />

effect of drop size on efficiency of splash-cup and<br />

springboard dispersal devices. Canadian Journal of<br />

Botany, 56, 127 128.<br />

Sawin, K. E. & Nurse, P. (1998). Regulation of cell<br />

polarity by microtubules in fission yeast. Journal of<br />

Cell Biology, 142, 457 471.<br />

Sawin, K. E. & Snaith, H. A. (2004). Role of microtubules<br />

and tea1p in establishment and maintenance of<br />

fission yeast cell polarity. Journal of Cell Science, 117,<br />

689 700.<br />

Scarborough, G. A. (1970). Sugar transport in Neurospora<br />

crassa. Journal of Biological Chemistry, 245, 1694 1698.<br />

Schaad, N. W., Frederick, R. D., Shaw, J., et al. (2003).<br />

Advances in molecular-based diagnostics in meeting<br />

crop biosecurity and phy<strong>to</strong>sanitary issues. Annual<br />

Review of Phy<strong>to</strong>pathology, 41, 305 324.<br />

Schardl, C. L. (1996). Epichloe species: fungal symbionts<br />

of grasses. Annual Review of Phy<strong>to</strong>pathology, 34,<br />

109 130.<br />

Schardl, C. L. & Clay, K. (1997). Evolution of mutualistic<br />

endophytes from plant pathogens. In The Mycota VB:<br />

Plant Relationships, ed. G. C. Carroll & P. Tudzynski.<br />

Berlin: Springer-Verlag, pp. 221 238.<br />

Schardl, C. L. & Moon, C. D. (2003). Process of species<br />

evolution in Epichloe/Neotyphodium endophytes of<br />

grasses. In Clavicipitalean <strong>Fungi</strong>, ed. J. F. White, C. W.<br />

Bacon, N. L. Hywel-Jones & J. W. Spatafora. New York:<br />

Marcel Dekker Inc., pp. 273 310.<br />

Scheffer, R. P. (1992). Ecological and evolutionary role<br />

of <strong>to</strong>xins from Alternaria species. In Alternaria Biology,<br />

Plant Diseases and Metabolites, ed. J. Chelkowski & A.<br />

Visconti. Amsterdam: Elsevier, pp. 101 122.<br />

Schekman, R. (1992). Genetic and biochemical analysis<br />

of vesicular traffic in yeast. Current Opinion in Cell<br />

Biology, 4, 587 592.<br />

Schell, W. A. (2003). Dematiaceous hyphomycetes. In<br />

Pathogenic <strong>Fungi</strong> in Humans and Animals, ed.<br />

D. H. Howard. New York: Marcel Dekker Inc.,<br />

pp. 565 636.<br />

Scherrer, S., de Vries, O. M. H., Dudler, R., Wessels,<br />

J. G. H. & Honegger, R. (2000). Interfacial self-assembly<br />

of fungal hydrophobins of the lichen-forming<br />

ascomycetes Xanthoria parietina and X. ectaneoides.<br />

Fungal Genetics and Biology, 30, 81 93.<br />

Schieber, E. & Zentmyer, G. A. (1984). Coffee rust in the<br />

Western hemisphere. Plant Disease, 68, 89 93.<br />

Schiltz, P. (1981). Downy mildew of <strong>to</strong>bacco. In The<br />

Downy Mildews, ed. D. M. Spencer. London: Academic<br />

Press, pp. 577 599.


792 REFERENCES<br />

Schimek, C., Eibel, P., Grolig, F., et al. (1999).<br />

Gravitropism in Phycomyces: a role for sedimenting<br />

protein crystals and floating lipid globules. Planta,<br />

210, 132 142.<br />

Schimek, C., Kleppe, K., Saleem, A.-R., et al. (2003).<br />

Sexual reactions in Mortierellales are mediated by the<br />

trisporic acid system. Mycological Research, 107,<br />

736 747.<br />

Schipper, M. A. A. (1978). On certain species of Mucor,<br />

with a key <strong>to</strong> all accepted species. Studies in Mycology,<br />

17, 1 52.<br />

Schipper, M. A. A. (1987). Mating ability and the species<br />

concept in the Zygomycetes. In Evolutionary Biology of<br />

the <strong>Fungi</strong>, ed. A. D. M. Rayner, C. M. Brasier & D.<br />

Moore. Cambridge: Cambridge University Press,<br />

pp. 261 269.<br />

Schipper, M. A. A. & Stalpers, J. A. (1980). Various aspects<br />

of the mating system in Mucorales. Persoonia, 11,<br />

53 63.<br />

Schippers, B. & van Eck, W. H. (1981). Formation and<br />

survival of chlamydospores in Fusarium. In Fusarium:<br />

Diseases, Biology and Taxonomy, ed. P. E. Nelson, T. A.<br />

Toussoun & R. J. Cook. London, PA: Pennsylvania<br />

State University Press, pp. 250 260.<br />

Schmatz, D. M., Romancheck, M. A., Pittarelli, L. A., et al.<br />

(1990). Treatment of Pneumocystis carinii pneumonia<br />

with 1,3,-b-glucan synthesis inhibi<strong>to</strong>rs. Proceedings of<br />

the National Academy of Sciences of the USA, 87,<br />

5950 5954.<br />

Schmidt, E. (1983). Spore germination of and carbohydrate<br />

colonization by Morchella esculenta at different<br />

soil temperatures. Mycologia, 75, 870 875.<br />

Schmidt, F. R. (2002). Beta-lactam antibiotics: aspects of<br />

manufacture and therapy. In The Mycota X: Industrial<br />

Applications, ed. H. D. Osiewacz. Berlin: Springer-<br />

Verlag, pp. 69 91.<br />

Schmidt, O. & Moreth-Kebernik, U. (1991). Monokaryon<br />

pairing of the dry rot fungus Serpula lacrymans.<br />

Mycological Research, 95, 1382 1386.<br />

Schmidt, R. A. (2003). Fusiform rust of Southern pines:<br />

a major success for forest disease management.<br />

Phy<strong>to</strong>pathology, 93, 1048 1051.<br />

Scholler, M., Hagedorn, G. & Rubner, A. (1999). A<br />

reevaluation of preda<strong>to</strong>ry orbiliaceous fungi. II. A<br />

new generic concept. Sydowia, 51, 89 113.<br />

Schoonbeek, H., del Sorbo, G. & de Waard, M. A. (2001).<br />

The ABC transporter BcatrB affects the sensitivity of<br />

Botrytis cinerea <strong>to</strong> the phy<strong>to</strong>alexin resvera<strong>to</strong>l and the<br />

fungicide fenpiclonil. Molecular Plant Microbe<br />

Interactions, 14, 562 571.<br />

Schreurs, W. J. A., Harold, R. L. & Harold, F. M. (1989).<br />

Chemotropism and branching as alternate responses<br />

of Achlya bisexualis. Journal of General Microbiology, 135,<br />

2519 2528.<br />

Schroeder, W. A. & Johnson, E. A. (1993). Antioxidant<br />

role of carotenoids in Phaffia rhodozyma. Journal of<br />

General Microbiology, 139, 907 912.<br />

Schroeder, W. A. & Johnson, E. A. (1995). Carotenoids<br />

protect Phaffia rhodozyma against singlet oxygen<br />

damage. Journal of Industrial Microbiology, 14,<br />

502 507.<br />

Schultz, M., Arendholz, W.-R. & Büdel, B. (2001). Origin<br />

and evolution of the lichenized ascomycete order<br />

Lichinales: monophyly and systematic relationships<br />

inferred from ascus, fruiting body and SSU rDNA<br />

evolution. Plant Biology, 3, 116 123.<br />

Schulz, T. R., Johns<strong>to</strong>n, W. J., Goloh, C. T. & Maguire,<br />

J. D. (1993). Control of ergot in Kentucky bluegrass<br />

seeds using fungicides. Plant Disease, 77, 685 687.<br />

Schuman, G. L. (1991). Plant Diseases: Their Biology and<br />

Social Impact. St. Paul: APS Press.<br />

Schüssler, A. & Kluge, M. (2001). Geosiphon pyriforme, an<br />

endocy<strong>to</strong>symbiosis between fungus and cyanobacteria,<br />

and its meaning as a model system for<br />

arbuscular mycorrhizal research. In The Mycota IX:<br />

Fungal Associations, ed. B. Hock. Berlin: Springer-<br />

Verlag, pp. 151 161.<br />

Schüssler, A., Schwarzott, D. & Walker, C. (2001). A new<br />

fungal phylum, the Glomeromycota: phylogeny and<br />

evolution. Mycological Research, 105, 1413 1421.<br />

Schwendener, S. (1867). Über die wahre Natur der<br />

Flechtengonidien. Verhandlungen der Schweizerischen<br />

Naturforschenden Gesellschaft, 51, 88 90.<br />

Schwinn, F. J. & Staub, T. (1995). Phenylamides and<br />

other fungicides against Oomycetes. In Modern<br />

Selective <strong>Fungi</strong>cides. Properties, Applications, Mechanisms<br />

of Action, 2nd edn, ed. H. Lyr. New York: Gustav<br />

Fischer Verlag, pp. 323 346.<br />

Scorzetti, G., Fell, J. W., Fonseca, A. & Statzell-Tallman,<br />

A. (2002). Systematics of basidiomyce<strong>to</strong>us yeasts: a<br />

comparison of large subunit D1/D2 and internal<br />

transcribed spacer rDNA regions. FEMS Yeast Research,<br />

2, 495 517.<br />

Scott, W. W. (1961). A monograph of the genus<br />

Aphanomyces. Virginia Agricultural Experiment Station<br />

Technical Bulletin, 151, 1 95.<br />

Scrimshaw, N. S. & Murray, E. B. (1995). Nutritional<br />

value and safety of ‘single cell protein’. In<br />

Biotechnology 9: Enzymes, Biomass, Food and Feed, ed. G.<br />

Reed & T. W. Nagodawithana. Weinheim, Germany:<br />

VCH, pp. 221 237.<br />

Scudamore, K. A. (1994). Aspergillus <strong>to</strong>xins in food and<br />

animal feedingstuffs. In The Genus Aspergillus. From<br />

Taxonomy and Genetics <strong>to</strong> Industrial Application, ed.


REFERENCES<br />

793<br />

K. A. Powell, A. Renwick & J. F. Peberdy. New York:<br />

Plenum Press, pp. 59 71.<br />

Scutt, C. P., Li, Y., Robertson, S. E., Willis, M. E. &<br />

Gilmartin, P. M. (1997). Sex determination in dioecious<br />

Silene latifolia. Effects of the Y chromosome and<br />

the parasitic smut fungus (Ustilago violacea) on gene<br />

expression during flower development. Plant<br />

Physiology, 114, 969 979.<br />

Seaward, M. R. D. (1993). Lichens and sulphur dioxide<br />

air pollution: field studies. Environmental Reviews, 1,<br />

73 91.<br />

Seaward, M. R. D. (1997). Urban deserts bloom: a lichen<br />

renaissance. Bibliotheca Lichenologica, 67, 297 309.<br />

Seaward, M. R. D. (2004). The use of lichens for environmental<br />

impact assessment. Symbiosis, 37, 293 305.<br />

Seaward, M. R. D. & Coppins, B. J. (2004). Lichens and<br />

hypertrophication. Bibliotheca Lichenologica, 88,<br />

561 572.<br />

Seaward, M. R. D. & Edwards, H. G. M. (1997). Biological<br />

origin of major chemical disturbances on ecclesiastical<br />

architecture studied by Fourier transform<br />

Raman spectroscopy. Journal of Raman Spectroscopy,<br />

28, 691 696.<br />

Seeger, M. & Payne, G. S. (1992). A role for clathrin in<br />

the sorting of vacuolar proteins in the Golgi complex<br />

of yeast. EMBO Journal, 11, 2811 2818.<br />

Seifert, K. A. (1983). Decay of wood by the<br />

Dacrymycetales. Mycologia, 75, 1011 1018.<br />

Seifert, K. A. (1985). A monograph of Stilbella and some<br />

allied hyphomycetes. Studies in Mycology, 27, 1 224.<br />

Seifert, K. A. (1993). Sapstain of commercial lumber by<br />

species of Ophios<strong>to</strong>ma and Cera<strong>to</strong>cystis. In Cera<strong>to</strong>cystis<br />

and Ophios<strong>to</strong>ma. Taxonomy, Ecology and Pathogenicity,<br />

ed. M. J. Wingfield, K. A. Seifert & J. F. Webber. St.<br />

Paul, USA: APS Press, pp. 141 151.<br />

Seifert, K. A. & Gams, W. (2001). The taxonomy of<br />

anamorphic fungi. In The Mycota VIIA: Systematics and<br />

Evolution, ed. D. J. McLaughlin, E. G. McLaughlin &<br />

P. A. Lemke. Berlin: Springer-Verlag, pp. 307 347.<br />

Seifert, K. A. & Samuels, G. J. (2000). How should we<br />

look at anamorphs? Studies in Mycology, 45, 5 18.<br />

Seifert, K. A., Wingfield, M. J. & Kendrick, W. B. (1993).<br />

A nomencla<strong>to</strong>r for described species of Cera<strong>to</strong>cystis,<br />

Ophios<strong>to</strong>ma, Cera<strong>to</strong>cystiopsis, Cera<strong>to</strong>s<strong>to</strong>mella and<br />

Sphaeronaemella. In Cera<strong>to</strong>cystis and Ophios<strong>to</strong>ma.<br />

Taxonomy, Ecology and Pathogenicity, ed. M. J.<br />

Wingfield, K. A. Seifert & J. F. Webber. St. Paul, USA:<br />

APS Press, pp. 269 287.<br />

Seiler, S., Nargang, F. E., Steinberg, G. & Schliwa, M.<br />

(1997). Kinesin is essential for cell morphogenesis<br />

and polarized secretion in Neurospora crassa. EMBO<br />

Journal, 16, 3025 3034.<br />

Sela-Buurlage, M. B., Budai-Hadrian, O., Pan, Q., et al.<br />

(2001). Genome-wide dissection of Fusarium resistance<br />

in <strong>to</strong>ma<strong>to</strong> reveals multiple complex loci.<br />

Molecular Genetics and Genomics, 265, 1104 1111.<br />

Sentandreu, R., Mormeneo, S. & Ruiz-Herrera, J. (1994).<br />

Biogenesis of the fungal cell wall. In The Mycota I:<br />

Growth, Differentiation and Sexuality, ed.<br />

J. G. H. Wessels & F. Meinhardt. Berlin:<br />

Springer-Verlag, pp. 111 124.<br />

Seppelt, R. D. (1995). Phy<strong>to</strong>geography of continental<br />

Antarctic lichens. Lichenologist, 27, 417 431.<br />

Setliff, E. C. (1988). Hyphal deterioration in Ganoderma<br />

applanatum. Mycologia, 80, 447 454.<br />

Seviour, R. J., Willing, R. R. & Chilvers, G. A. (1973).<br />

Basidiocarps associated with ericoid mycorrhizas.<br />

New Phy<strong>to</strong>logist, 72, 381 385.<br />

Seymour, R. L. (1970). The genus Saprolegnia. Nova<br />

Hedwigia, 19, 1 124.<br />

Seymour, R L. & Johnson, T. W. (1973). Saprolegniaceae:<br />

a keratinophilic Aphanomyces from soil. Mycologia, 65,<br />

1312 1318.<br />

Shah, P. A., Aebi, M. & Tuor, T. (2000). Drying and<br />

s<strong>to</strong>rage procedures for formulated and unformulated<br />

mycelia of the aphid-pathogenic fungus Erynia<br />

neoaphidis. Mycological Research, 104, 440 446.<br />

Shamoun, S. F. (2000). Application of biological control<br />

<strong>to</strong> vegetation management in forestry. In Proceedings<br />

of the X International Symposium on Biological Control of<br />

Weeds, ed. N. R. Spencer. Bozeman, USA: Montana<br />

State University, pp. 87 96.<br />

Shankar, M., Cowling, W. A. & Sweetingham, M. W.<br />

(1998). His<strong>to</strong>logical observations of latent infection<br />

and tissue colonisation by Diaporthe <strong>to</strong>xica in resistant<br />

and susceptible narrow-leafed lupins. Canadian<br />

Journal of Botany, 76, 1305 1316.<br />

Shapira, R., Choi, G. H., Hillman, B. I. & Nuss, D. L.<br />

(1991). The contribution of defective RNAs <strong>to</strong><br />

complexity of viral-encoded double-stranded RNA<br />

populations present in hypovirulent strains of the<br />

chestnut blight fungus, Cryphonectria parasitica. EMBO<br />

Journal, 10, 741 746.<br />

Sharland, P. & Rayner, A. D. M. (1986). Mycelial interactions<br />

in Daldinia concentrica. Transactions of the<br />

British Mycological Society, 86, 643 649.<br />

Sharma, R. & Cammack, R. H. (1976). Spore germination<br />

and taxonomy of Synchytrium endobioticum and S.<br />

succisae. Transactions of the British Mycological Society,<br />

66, 137 147.<br />

Sharma, R., Rajak, R. C. & Pandey, A. K. (2002). Teaching<br />

techniques for mycology: 19. A micro-dilution droptail<br />

method for isolating Onygenalean ascomycetes<br />

from hair baits. Mycologist, 16, 153 157.


794 REFERENCES<br />

Shat<strong>to</strong>ck, R. C. & Preece, T. F. (2000). Tranzschel revisited:<br />

modern studies of the relatedness of different<br />

rust fungi confirm his Law. Mycologist, 14, 113 117.<br />

Shat<strong>to</strong>ck, R. C., Tooley, P. W. & Fry, W. E. (1986a).<br />

Genetics of Phy<strong>to</strong>phthora infestans: characterization of<br />

single oospore cultures from A1 isolates induced <strong>to</strong><br />

self by intraspecific stimulation. Phy<strong>to</strong>pathology, 76,<br />

407 410.<br />

Shat<strong>to</strong>ck, R. C., Tooley, P. W. & Fry, W. E. (1986b).<br />

Genetics of Phy<strong>to</strong>phthora infestans: determination of<br />

recombination, segregation and selfing by isozyme<br />

analysis. Phy<strong>to</strong>pathology, 76, 410 413.<br />

Shaw, D. E. (1993). Honeybees collecting Neurospora<br />

spores from steamed Pinus logs in Queensland.<br />

Mycologist, 7, 182 185.<br />

Shaw, D. S. (1983). The cy<strong>to</strong>genetics and genetics of<br />

Phy<strong>to</strong>phthora. InPhy<strong>to</strong>phthora. Its Biology, Taxonomy,<br />

Ecology, and Pathology, ed. D. C. Erwin, S. Bartnicki-<br />

Garcia & P. H. Tsao. St. Paul: APS Press, pp. 81 94.<br />

Shaw, G. C. & Kile, G. A., eds. (1991). Armillaria Root<br />

Disease. Agricultural Handbook 691. Washing<strong>to</strong>n DC:<br />

USDA Forestry Service.<br />

Shaw, J. D., Cummings, K. B., Huyer, G., Michaelis, S. &<br />

Wendland, B. (2001). Yeast as a model system for<br />

studying endocy<strong>to</strong>sis. Experimental Cell Research, 271,<br />

1 9.<br />

Shearer, C. A. (1992). The role of woody debris. In The<br />

Ecology of Aquatic Hyphomycetes, ed. F. Bärlocher.<br />

Berlin: Springer-Verlag, pp. 77 98.<br />

Shelby, R. A. (1999). Toxicology of ergot alkaloids in<br />

agriculture. In Ergot. The Genus Claviceps, ed. V. Křen &<br />

L. Cvak. Amsterdam: Harwood Academic Publishers,<br />

pp. 46 51.<br />

Sheu, Y.-J. & Snyder, M. (2001). Control of cell polarity<br />

and shape. In The Mycota VIII: Biology of the Fungal Cell,<br />

ed. R. J. Howard & N. A. R. Gow. Berlin: Springer-<br />

Verlag, pp. 19 53.<br />

Shoemaker, R. A. (1955). Biology, cy<strong>to</strong>logy and taxonomy<br />

of Cochliobolus sativus. Canadian Journal of Botany,<br />

33, 562 576.<br />

Shoemaker, R. A. & Brun, H. (2001). The teleomorph of<br />

the weakly aggressive segregate of Lep<strong>to</strong>sphaeria<br />

maculans. Canadian Journal of Botany, 79, 412 419.<br />

Sholberg, P. L. & Gaudet, D. A. (1992). Grass as a source<br />

of inoculum for rot caused by Coprinus psychromorbidus.<br />

Canadian Journal of Plant Pathology, 14, 221 226.<br />

Shorrocks, B. & Charlesworth, P. (1982). A field study of<br />

the association between the stinkhorn Phallus impudicus<br />

Pers. and the British fungal-breeding Drosophila.<br />

Biological Journal of the Linnean Society, 17, 307 318.<br />

Shykoff, J. A. & Bucheli, E. (1995). Pollina<strong>to</strong>r visitation<br />

patterns, floral rewards and the probability of<br />

transmission of Microbotryum violaceum, a venereal<br />

disease of plants. Journal of Ecology, 83, 189 198.<br />

Shykoff, J. A. & Kaltz, O. (1998). Phenotypic changes in<br />

host plants diseased by Microbotryum violaceum:<br />

parasite manipulation, side effects, and trade-offs.<br />

International Journal of Plant Sciences, 159, 236 243.<br />

Siddiqui, Z. A. & Mahmood, I. (1996). Biological control<br />

of plant parasitic nema<strong>to</strong>des by fungi: a review.<br />

Bioresource Technology, 58, 229 239.<br />

Sidhu, G. & Person, C. (1972). Genetic control of<br />

virulence in Ustilago hordei. III. Identification of genes<br />

for host resistance and demonstration of gene-forgene<br />

relations. Canadian Journal of Genetics and<br />

Cy<strong>to</strong>logy, 14, 209 213.<br />

Sidney, E. (1846). Blights of the Wheat and Their Remedies.<br />

London: The Religious Tract Society.<br />

Siegel, M. R. & Bush, L. P. (1997). Toxin production in<br />

grass/endophyte associations. In The Mycota VA: Plant<br />

Relationships, ed. G. C. Carroll & P. Tudzynski. Berlin:<br />

Springer-Verlag, pp. 185 207.<br />

Siegel, M. R., Johnson, M. C., Varney, D. R., et al. (1984). A<br />

fungal endophyte in tall fescue: incidence and<br />

dissemination. Phy<strong>to</strong>pathology, 74, 932 937.<br />

Sietsma, J. H. & Wessels, J. G. H. (1990). Occurrence of<br />

glucosaminoglycan in the cell wall of<br />

Schizosaccharomyces pombe. Journal of General<br />

Microbiology, 136, 2261 2265.<br />

Sietsma, J. H. & Wessels, J. G. H. (1994). Apical wall<br />

biogenesis. In The Mycota I: Growth, Differentiation and<br />

Sexuality, ed. J. G. H. Wessels & F. Meinhardt. Berlin:<br />

Springer-Verlag, pp. 125 141.<br />

Sietsma, J. H., Rast, D. & Wessels, J. G. H. (1977). The<br />

effect of carbon dioxide on fruiting and on the<br />

degradation of a cell wall glucan in Schizophyllum<br />

commune. Journal of General Microbiology, 102,<br />

385 389.<br />

Sigler, L. (2003). Ascomycetes: the Onygenaceae and<br />

other fungi from the order Onygenales. In Pathogenic<br />

<strong>Fungi</strong> in Humans and Animals, 2nd edn, ed. D. H.<br />

Howard. New York: Marcel Dekker, pp. 195 236.<br />

Sigler, L., Flis, A. L. & Carmichael, J. W. (1998). The genus<br />

Uncinocarpus (Onygenaceae) and its synonym<br />

Brunneospora: new concepts, combinations and<br />

connections <strong>to</strong> anamorphs in Chrysosporium, and<br />

further evidence of its relationship with Coccidioides<br />

immitis. Canadian Journal of Botany, 76, 1624 1636.<br />

Silar, P. (2005). Peroxide accumulation and cell death<br />

in filamen<strong>to</strong>us fungi induced by contact with a<br />

contestant. Mycological Research, 109, 137 149.<br />

Silar, P., Lalucque, H. & Vierney, C. (2001). Cell<br />

degeneration in the model system Podospora anserina.<br />

Biogeron<strong>to</strong>logy, 2, 1 17.


REFERENCES<br />

795<br />

Silliker, M. E., Monroe, J. A. & Jorden, M. A. (1997).<br />

Evaluation of the efficiency of sexual reproduction<br />

in res<strong>to</strong>ring Podospora anserina mi<strong>to</strong>chondrial DNA <strong>to</strong><br />

wild-type. Current Genetics, 32, 281 286.<br />

Silva, M. C., Nicole, M., Guerra-Guimarães, L. &<br />

Rodrigues, C. J. (2002). Hypersensitive cell death and<br />

post-haus<strong>to</strong>rial defence responses arrest the orange<br />

rust (Hemileia vastatrix) growth in resistant coffee<br />

leaves. Physiological and Molecular Plant Pathology, 60,<br />

169 183.<br />

Silva-Hanlin, D. M. W. & Hanlin, R. T. (1999). Small<br />

subunit ribosomal RNA gene phylogeny of several<br />

loculoascomycetes and its taxonomic implications.<br />

Mycological Research, 103, 153 160.<br />

Silverman-Gavrila, L. B. & Lew, R. R. (2001). Regulation<br />

of the tip-high Ca 2þ gradient in growing hyphae of<br />

the fungus Neurospora crassa. European Journal of Cell<br />

Biology, 80, 379 390.<br />

Silverman-Gavrila, L. B. & Lew, R. R. (2002). An<br />

IP 3 -activated Ca 2þ channel regulates fungal tip<br />

growth. Journal of Cell Science, 115, 5013 5025.<br />

Simmons, E. G. (1969). Perfect states of Stemphylium.<br />

Mycologia, 61, 1 26.<br />

Simmons, E. G. (1986). Alternaria themes and variations.<br />

Mycotaxon, 25, 287 309.<br />

Simmons, E. G. (1992). Alternaria taxonomy: current<br />

status, viewpoint, challenge. In Alternaria Biology,<br />

Plant Diseases and Metabolites, ed. J. Chelkowski & A.<br />

Visconti. Amsterdam: Elsevier, pp. 1 35.<br />

Simmons, R. B. & Ahern, D. G. (1987). Cell wall ultrastructure<br />

and diazonium blue B reaction of<br />

Sporopachydermia quercuum, Bullera tsugae, and<br />

Malassezia spp. Mycologia, 79, 38 43.<br />

Simon, L., Bousquet, J., Lévesque, R. C. & Lalonde, M.<br />

(1993). Origin and diversification of endomycorrhizal<br />

fungi and coincidence with vascular land plants.<br />

Nature, 363, 67 69.<br />

Simons, M. D. (1985). Crown rust. In The Cereal Rusts 2:<br />

Diseases, Distribution, Epidemiology, and Control, ed. A. P.<br />

Roelfs & W. R. Bushnell. Orlando: Academic Press,<br />

pp. 131 172.<br />

Simpson, R. M., van Hekezen, R., van Lune, F., et al.<br />

(2001). Extracellular enzymes of Chondrostereum<br />

purpureum, causal fungus of silverleaf disease. New<br />

Zealand Plant Protection, 54, 202 208.<br />

Sims, K. P., Watling, R. & Jeffries, P. (1995). A revised key<br />

<strong>to</strong> the genus Scleroderma. Mycotaxon, 56, 403 420.<br />

Singer, R., ed. (1986). The Agaricales in Modern Taxonomy,<br />

4th edn. Königstein, Germany: Koeltz Scientific<br />

Books.<br />

Singh, J., Bech-Andersen, J., Elborne, S. A., et al.<br />

(1993). The search for wild dry rot fungus<br />

(Serpula lacrymans) in the Himalayas. Mycologist,<br />

7, 124 130.<br />

Singh, R. P., Huerta-Espino, J. & Roelfs, A. P. (2002). The<br />

wheat rusts. In Bread Wheat. Improvement and<br />

Production, ed. B. C. Curtis, S. Rajaram & H. Gómez<br />

Macpherson. Rome: Food and Agriculture<br />

Organization of the United Nations, pp. 227 249.<br />

Sipiczki, M. & Bozsik, A. (2000). The use of morphomutants<br />

<strong>to</strong> investigate septum formation and cell<br />

separation in Schizosaccharomyces pombe. Archives of<br />

Microbiology, 174, 386 392.<br />

Sipman, H. J. M. & Aptroot, A. (2001). Where are the<br />

missing lichens? Mycological Research, 105,<br />

1433 1439.<br />

Sivanesan, A. (1984). The Bitunicate Ascomycetes and their<br />

Anamorphs. Lehre, Germany: J. Cramer.<br />

Sivanesan, A. (1987). Graminicolous species of Bipolaris,<br />

Curvularia, Drechslera, Exserohilum and their teleomorphs.<br />

Mycological Papers, 158, 1 261.<br />

Sivichai, S., Hywel-Jones, N. & Somrithipol, S. (2000).<br />

Lignicolous freshwater Ascomycota from Thailand:<br />

Melanochaeta and Sporoschisma anamorphs. Mycological<br />

Research, 104, 478 485.<br />

Sjamsuridzal, W., Tajiri, Y., Nishida, H., et al. (1997).<br />

Evolutionary relationships of members of the genera<br />

Taphrina, Pro<strong>to</strong>myces, Schizosaccharomyces, and related<br />

taxa within the archiascomycetes: Integrated analysis<br />

of genotypic and phenotypic characters.<br />

Mycoscience, 38, 267 280.<br />

Sjollema, K. A., Dijksterhuis, J., Veenhuis, M. & Harder,<br />

W. (1993). An electron microscopical study of the<br />

infection of the nema<strong>to</strong>de Panagrellus redivivus by the<br />

endoparasitic fungus Verticillium balanoides.<br />

Mycological Research, 97, 479 484.<br />

Skou, J. P. (1972). Ascosphaerales. Friesia, 10, 1 24.<br />

Skou, J. P. (1975). Two new species of Ascosphaera and<br />

notes on the conidial state of Bettsia alvei. Friesia, 11,<br />

62 74.<br />

Skou, J. P. (1982). Ascosphaerales and their unique<br />

ascomata. Mycotaxon, 15, 487 499.<br />

Skou, J. P. (1988). More details in support of the class<br />

Ascosphaeromycetes. Mycotaxon, 31, 191 198.<br />

Skucas, G. P. (1967). Structure and composition of the<br />

resistant sporangial wall in the fungus Allomyces.<br />

American Journal of Botany, 54, 1152 1158.<br />

Skucas, G. P. (1968). Changes in wall and internal<br />

structure of Allomyces resistant sporangia during<br />

germination. American Journal of Botany, 55, 291 295.<br />

Slaninova, I., Kucsera, J. & Svoboda, A. (1999). Topology<br />

of microtubules and actin in the life cycle of<br />

Xanthophyllomyces dendrorhous (Phaffia rhodozyma).<br />

An<strong>to</strong>nie van Leeuwenhoek, 75, 361 368.


796 REFERENCES<br />

Sláviková, E. & Vadkertiová, R. (2003). The diversity of<br />

yeasts in the agricultural soil. Journal of Basic<br />

Microbiology, 43, 430 436.<br />

Slayman, C. L. (1987). The plasma membrane ATPase of<br />

Neurospora: a pro<strong>to</strong>n-pumping electroenzyme. Journal<br />

of Bioenergetics and Biomembranes, 19, 1 20.<br />

Slayman, C. L. & Slayman, C. W. (1974). Depolarization<br />

of the plasma membrane of Neurospora during active<br />

transport of glucose: evidence for a pro<strong>to</strong>n-dependent<br />

cotransport system. Proceedings of the National<br />

Academy of Sciences of the USA, 71, 1935 1939.<br />

Slippers, B., Coutinho, T. A., Wingfield, B. D. &<br />

Wingfield, M. J. (2003). A review of the genus<br />

Amylostereum and its association with woodwasps.<br />

South African Journal of Science, 99, 70 74.<br />

Slutsky, B., Staebell, M., Anderson, J., et al. (1987).<br />

‘White opaque transition’: a second high-frequency<br />

switching system in Candida albicans. Journal of<br />

Bacteriology, 169, 189 197.<br />

Smalley, E. B. & Guries, R. P. (1993). Breeding elms for<br />

resistance <strong>to</strong> Dutch elm disease. Annual Review of<br />

Phy<strong>to</strong>pathology, 31, 325 352.<br />

Smalley, E. B., Raffa, K. F., Proc<strong>to</strong>r, R. H. & Klepzig, K. D.<br />

(1993). Tree responses <strong>to</strong> infection by species of<br />

Ophios<strong>to</strong>ma and Cera<strong>to</strong>cystis. In Cera<strong>to</strong>cystis and<br />

Ophios<strong>to</strong>ma. Taxonomy, Ecology and Pathogenicity, ed.<br />

M. J. Wingfield, K. A. Seifert & J. F. Webber. St. Paul,<br />

USA: APS Press, pp. 207 17.<br />

Smart, C. D., Yuan, W., Foglia, R., et al. (2000).<br />

Cryphonectria hypovirus 3, a virus species in the<br />

family Hypoviridae with a single open reading<br />

frame. Virology, 265, 66 73.<br />

Smith, D. A. (1994). A local-oscilla<strong>to</strong>r theory of shuttle<br />

streaming in Physarum polycephalum. II. Phase control<br />

by cy<strong>to</strong>plasmic calcium. Pro<strong>to</strong>plasma, 177, 171 180.<br />

Smith, I. M., Dunez, J., Lelliott, R. A., Phillips, D. H. &<br />

Archer, S. A., eds. (1988). European Handbook of Plant<br />

Diseases. Oxford: Blackwell Scientific Publications.<br />

Smith, J. R. & Rubenstein, I. (1973). The development of<br />

‘senescence’ in Podospora anserina. Journal of General<br />

Microbiology, 76, 283 296.<br />

Smith, M. F. & Callaghan, A. A. (1987). Quantitative<br />

survey of Conidiobolus and Basidiobolus in soils and<br />

litter. Transactions of the British Mycological Society, 89,<br />

179 185.<br />

Smith, M. L., Bruhn, J. N. & Anderson, J. B. (1992). The<br />

fungus Armillaria bulbosa is among the largest and<br />

oldest living organisms. Nature, 356, 428 431.<br />

Smith, S. E. (1966). Physiology and ecology of orchid<br />

mycorrhizal fungi. New Phy<strong>to</strong>logist, 65, 488 499.<br />

Smith, S. E. (1967). Carbohydrate translocation in<br />

orchid mycorrhizas. New Phy<strong>to</strong>logist, 66, 371 378.<br />

Smith, S. E. & Read, D. J. (1997). Mycorrhizal Symbiosis,<br />

2nd edn. London: Academic Press.<br />

Smreciu, E. A. & Currah, R. S. (1989). Symbiotic germination<br />

of seeds of terrestrial orchids of North<br />

America and Europe. Lindleyana, 4, 6 15.<br />

Sneh, B., Burpee, L. & Ogoshi, A. (1991). Identification of<br />

Rhizoc<strong>to</strong>nia Species. St. Paul: APS Press.<br />

Sneh, B., Jabaji-Hare, S., Neate, S. M. & Dijst, G., eds.<br />

(1996). Rhizoc<strong>to</strong>nia Species: Taxonomy, Molecular Biology,<br />

Ecology, Pathology and Disease Control. Dordrecht:<br />

Kluwer.<br />

Snetselaar, K. M. & Mims, C. W. (1992). Sporidial fusion<br />

and infection of maize seedlings by the smut fungus<br />

Ustilago maydis. Mycologia, 84, 193 203.<br />

Snetselaar, K. M. & Mims, C. W. (1993). Infection of<br />

maize stigmas by Ustilago maydis: light and electron<br />

microscopy. Phy<strong>to</strong>pathology, 83, 843 850.<br />

Snetselaar, K. M. & Mims, C. W. (1994). Light and<br />

electron microscopy of Ustilago maydis hyphae in<br />

maize. Mycological Research, 98, 347 355.<br />

Snetselaar, K. M., Bölker, M. & Kahmann, R. (1996).<br />

Ustilago maydis mating hyphae orient their growth<br />

<strong>to</strong>ward pheromone sources. Fungal Genetics and<br />

Biology, 20, 299 312.<br />

Snider, P. J. (1959). Stages of development in rhizomorphic<br />

thalli of Armillaria mellea. Mycologia, 51,<br />

693 707.<br />

Snider, P. J. (1965). Incompatibility and nuclear migration.<br />

In Incompatibility in <strong>Fungi</strong>, ed. K. Esser & J. R.<br />

Raper. Berlin: Springer-Verlag, pp. 52 70.<br />

Snider, P. J. (1968). Nuclear movements in Schizophyllum.<br />

Symposia of the Society for Experimental Biology, 22,<br />

261 283.<br />

Soares, M., Christen, P., Pandey, A. & Soccol, C. R. (2000).<br />

Fruity flavour production by Cera<strong>to</strong>cystis fimbriata<br />

grown on coffee husk in solid-state fermentation.<br />

Process Biochemistry, 35, 857 861.<br />

Söderhäll, K., Dick, M. W., Clark, G., Fürst, M. &<br />

Constantinescu, O. (1991). Isolation of Saprolegnia<br />

parasitica from the crayfish Astacus lep<strong>to</strong>dactylus.<br />

Aquaculture, 92, 121 125.<br />

Soll, D. R. (2002). Molecular biology of switching in<br />

Candida. InFungal Pathogenesis. Principles and Clinical<br />

Applications, ed. R. A. Calderone & R. L. Cihlar. New<br />

York: Marcel Dekker, pp. 161 82.<br />

Soll, D. R. (2003). Mating-type locus homozygosis,<br />

phenotypic switching and mating: a unique<br />

sequence of dependencies in Candida albicans.<br />

BioEssays, 26, 10 20.<br />

Soll, D. R., Lockhart, S. R. & Zhao, R. (2003). Relationship<br />

between switching and mating in Candida albicans.<br />

Eukaryotic Cell, 2, 390 397.


REFERENCES<br />

797<br />

Solla, A. & Gil, L. (2002). Vessel diameter as a fac<strong>to</strong>r in<br />

resistance of Ulmus minor <strong>to</strong> Ophios<strong>to</strong>ma novo-ulmi.<br />

Forest Pathology, 32, 123 134.<br />

Solomon, J. M., Rupper, A., Cardelli, J. A. & Isberg, R. R.<br />

(2000). Intracellular growth of Legionella pneumophila<br />

in Dictyostelium discoideum, a system for genetic<br />

analysis of host pathogen interactions. Infection and<br />

Immunity, 68, 2939 2947.<br />

Somers, E. & Horsfall, J. G. (1966). The water content of<br />

powdery mildew conidia. Phy<strong>to</strong>pathology, 56,<br />

1031 1035.<br />

Sonnenberg, A. S. M. (2000). Genetics and breeding of<br />

Agaricus bisporus. InScience and Cultivation of Edible<br />

<strong>Fungi</strong>, ed. L. J. L. D. van Griensven. Rotterdam: A.<br />

Balkema, pp. 25 39.<br />

Sorrell, T. C., Chen, S. C. A., Ruma, P., et al. (1996).<br />

Concordance of clinical and environmental isolates<br />

of Cryp<strong>to</strong>coccus neoformans var. gattii by random<br />

amplification of polymorphic DNA analysis and PCR<br />

fingerprinting. Journal of Clinical Microbiology, 34,<br />

1253 1260.<br />

Soylu, S., Keshavarzi, M., Brown, I. & Mansfield, J. W.<br />

(2003). Ultrastructural characterisation of interactions<br />

between Arabidopsis thaliana and Albugo candida.<br />

Physiological and Molecular Plant Pathology, 63,<br />

201 211.<br />

Sparrow, F. K. (1960). Aquatic Phycomycetes, 2nd edn. Ann<br />

Arbor: University of Michigan Press.<br />

Sparrow, F. K. (1973). Chytridiomycetes.<br />

Hyphochytridiomycetes. In The <strong>Fungi</strong>: An Advanced<br />

Treatise IVB, ed. G. C. Ainsworth, F. K. Sparrow &<br />

A. S. Sussman. New York: Academic Press,<br />

pp. 85 110.<br />

Spatafora, J. W. (1995). Ascomal evolution of filamen<strong>to</strong>us<br />

ascomycetes: evidence from molecular data.<br />

Canadian Journal of Botany, 73, S811 15.<br />

Spatafora, J. W. & Blackwell, M. (1993). Molecular<br />

systematics of unitunicate perithecial ascomycetes.<br />

I. The Clavicipitales Hypocreales connection.<br />

Mycologia, 85, 912 922.<br />

Spatafora, J. W. & Blackwell, M. (1994). The polyphyletic<br />

origins of ophios<strong>to</strong>ma<strong>to</strong>id fungi. Mycological Research,<br />

98, 1 9.<br />

Speare, R. & Thomas, A. D. (1985). Kangaroos and<br />

wallabies as carriers of Basidiobolus hap<strong>to</strong>sporus.<br />

Australian Veterinary Journal, 62, 209 210.<br />

Spellig, T., Bölker, M., Lottspeich, F., Frank, R. W. &<br />

Kahmann, R. (1994). Pheromones trigger filamen<strong>to</strong>us<br />

growth in Ustilago maydis. EMBO Journal, 13,<br />

1620 1627.<br />

Spencer, D. M., ed. (1978). The Powdery Mildews. London:<br />

Academic Press.<br />

Spencer, D. M., ed. (1981). The Downy Mildews. London:<br />

Academic Press.<br />

Spencer, J. F. T. & Spencer, D. M., eds. (1990). Yeast<br />

Technology. Berlin: Springer-Verlag.<br />

Spencer, J. F. T., Ragout de Spencer, A. L. & Laluce, C.<br />

(2002). Non-conventional yeasts. Applied Microbiology<br />

and Biotechnology, 58, 147 156.<br />

Spencer, M. A., Vick, M. C. & Dick, M. W. (2002). Revision<br />

of Aplanopsis, Pythiopsis and ‘subcentric’ Achlya species<br />

(Saprolegniomycetidae) using 18S rDNA and<br />

morphological data. Mycological Research, 106,<br />

549 560.<br />

Spencer-Phillips, P. T. N. (1997). Function of fungal<br />

haus<strong>to</strong>ria in epiphytic and endophytic infections.<br />

Advances in Botanical Research, 24, 309 333.<br />

Spicher, G. & Brümmer, J.-M. (1995). Baked goods. In<br />

Biotechnology 9: Enzymes, Biomass, Food and Feed, ed. G.<br />

Reed & T. W. Nagodawithana. Weinheim, Germany:<br />

VCH, pp. 241 319.<br />

Spiegel, F. W. (1984). Pro<strong>to</strong>stelium nocturnum, a new,<br />

minute, ballis<strong>to</strong>sporous pro<strong>to</strong>stelid. Mycologia, 76,<br />

443 447.<br />

Spiegel, F. W. (1990). Phylum plasmodial slime molds,<br />

class Pro<strong>to</strong>stelida. In Handbook of Pro<strong>to</strong>ctista, ed. L.<br />

Margulis, J. O. Corliss, M. Melkonian & D. J. Chapman.<br />

Bos<strong>to</strong>n: Jones & Bartlett, pp. 484 497.<br />

Spiegel, F. W. (1991). A proposed phylogeny of the<br />

flagellated pro<strong>to</strong>stelids. BioSystems, 25, 113 120.<br />

Spiegel, F. W., Olive, L. S. & Brown, R. M. (1979). Roles of<br />

actin during sporocarp culmination in the simple<br />

myce<strong>to</strong>zoan Planopro<strong>to</strong>stelium aurantium. Proceedings of<br />

the National Academy of Sciences of the USA, 76,<br />

2335 2339.<br />

Spielman, L. J., Drenth, A., Davidse, L. C., et al. (1991). A<br />

second world-wide migration and population<br />

displacement of Phy<strong>to</strong>phthora infestans? Plant<br />

Pathology, 40, 422 430.<br />

Spiers, A. G. & Hopcroft, D. H. (1994). Comparative<br />

studies of the poplar rusts Melampsora medusae, M.<br />

larici-populina and their interspecific hybrid M.<br />

medusae-populina. Mycological Research, 98, 889 903.<br />

Spil<strong>to</strong>ir, C. F. (1955). Life cycle of Ascosphaera apis<br />

(Pericystis apis). American Journal of Botany, 42,<br />

501 508.<br />

Sprey, B. (1988). Cellular and extracellular localization<br />

of endocellulase in Trichoderma reesei. FEMS<br />

Microbiology Letters, 55, 283 294.<br />

Springer, M. L. & Yanofsky, C. (1989). A morphological<br />

and genetic analysis of conidiophore development in<br />

Neurospora crassa. Genes and Development, 3, 559 571.<br />

Sridhar, K. R. & Bärlocher, F. (1992a). Aquatic hyphomycetes<br />

in spruce roots. Mycologia, 84, 580 584.


798 REFERENCES<br />

Sridhar, K. R. & Bärlocher, F. (1992b). Endophytic<br />

aquatic hyphomycetes of roots of spruce, birch and<br />

maple. Mycological Research, 96, 305 308.<br />

Srinivasan, M. C., Narasimhan, M. J. & Thirumalachar,<br />

M. J. (1964). Artificial culture of En<strong>to</strong>mophthora muscae<br />

and morphological aspects for differentiation of the<br />

genera En<strong>to</strong>mophthora and Conidiobolus. Mycologia, 56,<br />

683 691.<br />

Srinivasan, S., Vargas, M. M. & Roberson, R. W. (1996).<br />

Functional, organizational, and biochemical analysis<br />

of actin in hyphal tip cells of Allomyces macrogynus.<br />

Mycologia, 88, 57 70.<br />

Staats, M., van Baarlen, P. & van Kan, J. A. L. (2005).<br />

Molecular phylogeny of the plant pathogenic<br />

genus Botrytis and the evolution of host<br />

specificity. Molecular Biology and Evolution,<br />

22, 333 346.<br />

Staben, C. & Yanofsky, C. (1990). Neurospora crassa a<br />

mating-type region. Proceedings of the National Academy<br />

of Sciences of the USA, 87, 4917 4921.<br />

Stadler, M., Anke, H. & Sterner, O. (1993). Linoleic<br />

acid the nematicidal principle of several<br />

nema<strong>to</strong>phagous fungi and its production in trapforming<br />

submerged cultures. Archives of Microbiology,<br />

160, 401 405.<br />

Stahmann, K.-P. (1997). Vitamins, amino acids. In<br />

Fungal Biotechnology, ed. T. Anke. Weinheim,<br />

Germany: Chapman & Hall, pp. 81 90.<br />

Stahmann, K.-P., Schimz, K.-L. & Sahm, H. (1993).<br />

Purification and characterization of four extracellular<br />

1,3-b-glucanases of Botrytis cinerea. Journal of<br />

General Microbiology, 139, 2833 2840.<br />

Stahmann, K.-P., Revuelta, J. L. & Seulberger, H. (2000).<br />

Three biotechnological processes using Ashbya gossypii,<br />

Candida famata, orBacillus subtilis compete with<br />

chemical riboflavin production. Applied Microbiology<br />

and Biotechnology, 53, 509 516.<br />

Staib, P., Wirsching, S., Strauss, A. & Morschhauser, J.<br />

(2001). Gene regulation and host adaptation<br />

mechanisms in Candida albicans. International Journal<br />

of Medical Microbiology, 291, 183 188.<br />

Stakman, E. C. & Christensen, C. M. (1946). Aerobiology<br />

in relation <strong>to</strong> plant disease. Botanical Review, 12,<br />

205 253.<br />

Stakman, E. C. & Levine, M. N. (1922). The determination<br />

of biologic forms of Puccinia graminis on Triticum<br />

spp. Minnesota University Agricultural Experiment Station<br />

Technical Bulletin, 8.<br />

Stalpers, J. A. (1974). Spiniger, a new genus for the<br />

imperfect states of basidiomycetes. Proceedings of the<br />

Koninklijke Nederlandse Akademie van Wetenschappen<br />

Series C, 77, 402 407.<br />

Stamps, D. J., Waterhouse, G. M., Newhook, F. J. & Hall,<br />

G. S. (1990). Revised tabular key <strong>to</strong> the species of<br />

Phy<strong>to</strong>phthora. Mycological Papers, 162, 1 28.<br />

Stanghellini, M. E. & Hancock, J. G. (1971). Radial extent<br />

of the bean spermosphere and its relation <strong>to</strong> the<br />

behaviour of Pythium ultimum. Phy<strong>to</strong>pathology, 61,<br />

165 168.<br />

Stanier, R. Y. (1942). The culture and nutrient requirements<br />

of a chytridiaceous fungus. Journal of<br />

Bacteriology, 43, 499 520.<br />

Stankis, M. M., Specht, C. A. & Giasson, L. (1990). Sexual<br />

incompatibility in Schizophyllum commune: from classical<br />

genetics <strong>to</strong> a molecular view. Seminars in<br />

Developmental Biology, 1, 195 206.<br />

Staples, R. C. (2000). Research on the rust fungi during<br />

the twentieth century. Annual Review of<br />

Phy<strong>to</strong>pathology, 38, 49 69.<br />

Starks, P. T., Blackie, C. A. & Seeley, T. D. (2000).<br />

Fever in honeybee colonies. Naturwissenschaften,<br />

87, 229 231.<br />

Starmer, W. T., Ganter, P. F., Aberdeen, V.,<br />

Lachance, M.–A. & Phaff, H. J. (1987). The ecological<br />

role of killer yeasts in natural communities of yeasts.<br />

Canadian Journal of Microbiology, 33, 783 796.<br />

Statzell-Tallman, A. & Fell, J. W. (1998). Sporidiobolus<br />

Nyland. In The Yeasts: A Taxonomic Study, 4th edn, ed.<br />

C. P. Kurtzman & J. W. Fell. Amsterdam: Elsevier,<br />

pp. 693 699.<br />

Staub, T. (1991). <strong>Fungi</strong>cide resistance: practical experience<br />

with anti-resistance strategies and the role of<br />

integrated use. Annual Review of Phy<strong>to</strong>pathology, 29,<br />

421 442.<br />

Steinberg, G. (1998). Organelle transport and molecular<br />

mo<strong>to</strong>rs in fungi. Fungal Genetics and Biology, 24,<br />

161 177.<br />

Steinberg, G. (2000). The cellular roles of molecular<br />

mo<strong>to</strong>rs in fungi. Trends in Microbiology, 8, 162 168.<br />

Steinberg, G. & Fuchs, U. (2004). The role of microtubules<br />

in cellular organization and endocy<strong>to</strong>sis in<br />

the plant pathogen Ustilago maydis. Journal of<br />

Microscopy, 214, 114 123.<br />

Steinberg, G., Schliwa, M., Lehmler, C., et al. (1998).<br />

Kinesin from the plant pathogenic Ustilago maydis<br />

is involved in vacuole formation and<br />

cy<strong>to</strong>plasmic migration. Journal of Cell Science,<br />

111, 2235 2246.<br />

Steinert, M. & Heuner, K. (2005). Dictyostelium as host<br />

model for pathogenesis. Cellular Microbiology, 7,<br />

307 314.<br />

Stenroos, S. K. & DePriest, P. T. (1998). SSU rDNA<br />

phylogeny of cladoniiform lichens. American Journal<br />

of Botany, 85, 1548 1559.


REFERENCES<br />

799<br />

Stensrud, Ø., Hywel-Jones, N. L. & Schumacher, T.<br />

(2005). Towards a phylogenetic classification of<br />

Cordyceps: ITS nrDNA sequence data confirm divergent<br />

lineages and paraphyly. Mycological Research,<br />

109, 41 56.<br />

Stensvand, A., Eikemo, H., Gadoury, D. M. & Seem, R. C.<br />

(2005). Use of a rainfall frequency threshold <strong>to</strong> adjust<br />

a degree-day model of ascospore maturity of Venturia<br />

inaequalis. Plant Disease, 89, 198 202.<br />

Stephenson, S. L. & Stempen, H. (1994). Myxomycetes. A<br />

Handbook of Slime Molds. Portland, Oregon: Timber<br />

Press.<br />

Stergiopoulos, I., Gielkens, M. M., Goodall, S. D.,<br />

Venema, K. & de Waard, M. A. (2002). Molecular<br />

cloning and characterisation of three new ATPbinding<br />

cassette transporter genes from the wheat<br />

pathogen Mycosphaerella graminicola. Gene, 289,<br />

141 149.<br />

Sternberg, J. A., Geffken, D., Adams, J. B., et al. (2001).<br />

Famoxadone: the discovery and optimisation of a<br />

new agricultural fungicide. Pest Management Science,<br />

57, 143 152.<br />

Stevens, R. B., ed. (1974). Mycology Guidebook. Seattle:<br />

University of Washing<strong>to</strong>n Press.<br />

Stiles, C. M. & Glawe, D. A. (1989). Colonization of<br />

soybean roots by fungi isolated from cysts of<br />

Heterodera glycines. Mycologia, 81, 797 799.<br />

S<strong>to</strong>ll, M., Piepenbring, M., Begerow, D. &<br />

Oberwinkler, F. (2003). Molecular phylogeny of<br />

Ustilago and Sporisorium species (Basidiomycota,<br />

Ustilaginales) based on internal transcribed spacer<br />

(ITS) sequences. Canadian Journal of Botany, 81,<br />

976 984.<br />

Straatsma, G. & Samson, R. A. (1993). Taxonomy of<br />

Scytalidium thermophilum, an important thermophilic<br />

fungus in mushroom compost. Mycological Research,<br />

97, 321 328.<br />

Straatsma, G., Samson, R. A., Olijinsma, T. W., et al.<br />

(1994a). Ecology of thermophilic fungi in mushroom<br />

compost, with emphasis on Scytalidium thermophilum<br />

and growth stimulation of Agaricus bisporus mycelium.<br />

Applied and Environmental Microbiology, 60,<br />

454 458.<br />

Straatsma, G., Olijinsma, T. W., Gerrits, J. P. G., et al.<br />

(1994b). Inoculation of Scytalidium thermophilum in<br />

but<strong>to</strong>n mushroom compost and its effect on yield.<br />

Applied and Environmental Microbiology, 60,<br />

3049 3054.<br />

Straatsma, G., Samson, R. A., Olijinsma, T. W., et al.<br />

(1995). Bioconversion of cereal straw in<strong>to</strong> mushroom<br />

compost. Canadian Journal of Botany, 73,<br />

S1019 S1024.<br />

Strickmann, E. & Chadefaud, M. (1961). Recherches sur<br />

les asques et sur les périthèces des Nectria et<br />

réflexions sur l’évolution des Ascomycètes. Revue<br />

Générale de Botanique, 68, 725 770.<br />

S<strong>to</strong>ckdale, P. M. (1968). Sexual stimulation between<br />

Arthroderma simii S<strong>to</strong>ckdale, Mackenzie and<br />

Austwick and related species. Sabouraudia, 6,<br />

176 181.<br />

S<strong>to</strong>ev, S. D. (1998). The role of ochra<strong>to</strong>xin A as a possible<br />

cause of Balkan endemic nephropathy and its risk<br />

evaluation. Veterinary and Human Toxicology, 40,<br />

352 360.<br />

S<strong>to</strong>rck, R. & Morrill, R. C. (1977). Nuclei, nucleic acids<br />

and protein in sporangiospores of Mucor bacilliformis.<br />

Mycologia, 69, 1031 1041.<br />

S<strong>to</strong>rsberg, J., Schulz, H. & Keller, E. R. J. (2003).<br />

Chemotaxonomic classification of some Allium wild<br />

species on the basis of their volatile sulphur<br />

compounds. Journal of Applied Botany Angewandte<br />

Botanik, 77, 160 162.<br />

Strassmann, J. E., Zhu, Y. & Queller, D. C. (2000).<br />

Altruism and social cheating in the social amoeba<br />

Dictyostelium discoideum. Nature, 408, 965 967.<br />

Strathern, J. N. & Herskowitz, I. (1979). Asymmetry and<br />

directionality in production of new cell types during<br />

clonal growth: the switching pattern of homothallic<br />

yeast. Cell, 17, 371 381.<br />

Stringer, J. R. (1996). Pneumocystis carinii: What is it,<br />

exactly? Clinical Microbiology Reviews, 9, 489 498.<br />

Stringer, J. R. (2002). Pneumocystis. International Journal<br />

of Medical Microbiology, 292, 391 404.<br />

Stringer, J. R., Beard, C. B., Miller, R. F. & Wakefield, A. E.<br />

(2002). A new name (Pneumocystis jiroveci) for pneumocystis<br />

from humans. Emerging Infectious Diseases, 8,<br />

891 896.<br />

Stringer, M. A., Dean, R. A., Sewall, T. C. & Timberlake,<br />

W. E. (1991). Rodletness, a new Aspergillus developmental<br />

mutant induced by directed gene inactivation.<br />

Genes and Development, 5, 1161 1171.<br />

Struck, C., Hahn, M. & Mendgen, K. (1996). Plasma<br />

membrane H þ -ATPase activity in spores, germ tubes,<br />

and haus<strong>to</strong>ria of the rust fungus Uromyces viciae-fabae.<br />

Fungal Genetics and Biology, 20, 30 35.<br />

Strunz, G. M., Bethell, R., Dumas, M. T. & Boyonoski, N.<br />

(1997). On a new synthesis of sterpurene and the<br />

bioactivity of some related Chondrostereum purpureum<br />

sesquiterpene metabolites. Canadian Journal of<br />

Chemistry, 75, 742 753.<br />

Stubblefield, S. P. & Taylor, T. N. (1983). Studies of<br />

paleozoic fungi. I. The structure and organization of<br />

Traquairia (Ascomycota). American Journal of Botany,<br />

70, 387 399.


800 REFERENCES<br />

Stubblefield, S. P., Taylor, T. N., Miller, C. E. & Cole, G. T.<br />

(1983). Studies of carboniferous fungi. II. The<br />

structure and organization of Myocarpon,<br />

Sporocarpon, Dubiocarpon and Coleocarpon<br />

(Ascomycotina). American Journal of Botany, 70,<br />

1482 1498.<br />

Stubbs, R. W. (1985). Stripe rust. In The Cereal Rusts 2:<br />

Diseases, Distribution, Epidemiology, and Control, ed. A. P.<br />

Roelfs & W. R. Bushnell. Orlando: Academic Press,<br />

pp. 61 101.<br />

Stuteville, D. L. (1981). Downy mildew of forage<br />

legumes. In The Downy Mildews, ed. D. M. Spencer.<br />

London: Academic Press, pp. 355 66.<br />

Suarit, R., Gopal, P. K. & Shepherd, M. G. (1988).<br />

Evidence for a glycosidic linkage between chitin and<br />

glucan in the cell wall of Candida albicans. Journal of<br />

General Microbiology, 134, 1723 1730.<br />

Suberkropp, K. & Klug, M. J. (1976). <strong>Fungi</strong> and bacteria<br />

associated with dead leaves during processing in a<br />

woodland stream. Ecology, 57, 707 719.<br />

Suberkropp, K. & Klug, M. J. (1980). The maceration of<br />

deciduous leaf litter by aquatic hyphomycetes.<br />

Canadian Journal of Botany, 58, 1025 1031.<br />

Subramanian, C. V. (1971). The phialide. In Taxonomy of<br />

<strong>Fungi</strong> Imperfecti, ed. B. Kendrick. Toron<strong>to</strong>: Toron<strong>to</strong><br />

University Press, pp. 93 119.<br />

Sugita, T., Su<strong>to</strong>, H., Unno, T., et al. (2001). Molecular<br />

analysis of Malassezia microflora on the skin of<br />

a<strong>to</strong>pic dermatitis patients and healthy subjects.<br />

Journal of Clinical Microbiology, 39, 3486 3490.<br />

Sugiyama, J., ed. (1987). Pleomorphic <strong>Fungi</strong>: The Diversity<br />

and its Taxonomic Implications. Tokyo: Kodansha.<br />

Sugiyama, J., Summerbell, R. C. & Mikawa, T. (2002).<br />

Molecular phylogeny of onygenalean fungi based on<br />

small subunit (SSU) and large subunit (LSU) ribosomal<br />

DNA sequences. Studies in Mycology, 47, 5 23.<br />

Summerbell, R. C. (2000). Form and function in the<br />

evolution of derma<strong>to</strong>phytes. In Biology of<br />

Derma<strong>to</strong>phytes and Other Keratinolytic <strong>Fungi</strong>, ed.<br />

R. K. S. Kushwara & J. Guarro. Bilbao: Revista<br />

Iberoamericana de Micología, pp. 30 43.<br />

Summerbell, R. (2003). Ascomycetes: Aspergillus,<br />

Fusarium, Sporothrix, Piedraia, and their relatives. In<br />

Pathogenic <strong>Fungi</strong> in Humans and Animals, 2nd edn, ed.<br />

D. H. Howard. New York: Marcel Dekker,<br />

pp. 237 498.<br />

Summerbell, R. C., Kane, J., Krajden, S. & Duke, E. E.<br />

(1993). Medically important Sporothrix species and<br />

related ophios<strong>to</strong>ma<strong>to</strong>id fungi. In Cera<strong>to</strong>cystis and<br />

Ophios<strong>to</strong>ma. Taxonomy, Ecology and Pathogenicity, ed.<br />

M. J. Wingfield, K. A. Seifert & J. F. Webber. St. Paul:<br />

APS Press, pp. 185 92.<br />

Sun, N. C. & Bowen, C. C. (1972). Ultrastructural studies<br />

of nuclear division in Basidiobolus ranarum Eidam.<br />

Caryologia, 25, 243 247.<br />

Sundheim, L. (1982). Control of cucumber powdery<br />

mildew by the hyperparasite Ampelomyces quisqualis<br />

and fungicides. Plant Pathology, 31, 209 214.<br />

Sussman, A. S. (1968). Longevity and survivability of<br />

fungi. In The <strong>Fungi</strong>. An Advanced Treatise III: The Fungal<br />

Population, ed. G. C. Ainsworth & A. S. Sussman. New<br />

York Academic Press, pp. 447 86.<br />

Sussman, A. S. & Halvorson, H. O. (1966). Spores:<br />

Their Dormancy and Germination. New York: Harper &<br />

Row.<br />

Sutherland, M. L. & Brasier, C. M. (1995). Effect of d-<br />

fac<strong>to</strong>rs on in vitro cera<strong>to</strong>-ulmin production by the<br />

Dutch-elm disease pathogen Ophios<strong>to</strong>ma novo-ulmi.<br />

Mycological Research, 99, 1211 1217.<br />

Sutter, R. P. (1987). Sexual development. In Phycomyces,<br />

ed. E. Cerdá-Olmedo & E. D. Lipson. Cold Spring<br />

Harbor, New York: Cold Spring Harbor Labora<strong>to</strong>ry,<br />

pp. 317 36.<br />

Sut<strong>to</strong>n, B. C. (1980). The Coelomycetes. <strong>Fungi</strong> Imperfecti with<br />

Pycnidia, Acervuli and Stromata. Kew, UK:<br />

Commonwealth Mycological Institute.<br />

Sut<strong>to</strong>n, B. C. (1986). Presidential address.<br />

Improvisations on conidial themes. Transactions of the<br />

British Mycological Society, 86, 1 38.<br />

Sut<strong>to</strong>n, B. C. (1992). The genus Glomerella and its<br />

anamorph Colle<strong>to</strong>trichum. InColle<strong>to</strong>trichum: Biology,<br />

Pathology and Control, ed. J. A. Bailey & M. J. Jeger.<br />

Wallingford, UK: CAB International, pp. 1 26.<br />

Sut<strong>to</strong>n, D. K., MacHardy, W. E. & Lord, W. G. (2000).<br />

Effects of shredding or treating apple leaf litter with<br />

urea on ascospore dose of Venturia inaequalis and<br />

disease buildup. Plant Disease, 84, 1319 1326.<br />

Sut<strong>to</strong>n, P. N., Henry, M. J. & Hall, J. L. (1999). Glucose,<br />

and not sucrose, is transported from wheat <strong>to</strong> wheat<br />

powdery mildew. Planta, 208, 426 430.<br />

Suyan<strong>to</strong>, Ohtsuki, T., Yazaki, S., Ui, S. &<br />

Mimura, A. (2003). Isolation of a novel thermophilic<br />

fungus, Chae<strong>to</strong>mium sp. nov. MS-017 and description<br />

of its palm-oil fiber-decomposing<br />

properties. Applied Microbiology and Biotechnology, 60,<br />

581 587.<br />

Suzuki, M. & Nakase, T. (1999). A phylogenetic study of<br />

ubiquinone Q-8 species of the genera Candida, Pichia,<br />

and Citeromyces based on 18S ribosomal DNA<br />

sequence divergence. Journal of General and Applied<br />

Microbiology, 45, 239 246.<br />

Swann, E. C. & Taylor, J. W. (1993). Higher taxa of<br />

basidiomycetes: an 18S rRNA gene perspective.<br />

Mycologia, 85, 923 936.


REFERENCES<br />

801<br />

Swann, E. C. & Taylor, J. W. (1995). Phylogenetic<br />

perspectives on basidiomycete systematics: evidence<br />

from the 18S rRNA gene. Canadian Journal of Botany,<br />

73, S862 S868.<br />

Swann, E. C., Frieders, E. M. & McLaughlin, D. J. (1999).<br />

Microbotryum, Kriegeria and the changing paradigm<br />

in basidiomycete classification. Mycologia, 91, 51 66.<br />

Swann, E. C., Frieders, E. M. & McLaughlin, D. J. (2001).<br />

Urediniomycetes. In The Mycota VIIB: Systematics and<br />

Evolution, ed. D. J. McLaughlin, E. G. McLaughlin &<br />

P. A. Lemke. Berlin: Springer-Verlag, pp. 37 56.<br />

Swanson, J. A. & Taylor, D. L. (1982). Local and spatially<br />

coordinated movements in Dictyostelium discoideum<br />

amoebae during chemotaxis. Cell, 28, 225 232.<br />

Swedjemark, G. & Stenlid, J. (1993). Population<br />

dynamics of the root rot fungus Heterobasidion<br />

annosum following thinning of Picea abies. Oikos, 66,<br />

247 254.<br />

Sykes, E. E. & Porter, D. (1980). Infection and development<br />

of the obligate parasite Catenaria allomycis on<br />

Allomyces arbuscula. Mycologia, 72, 288 300.<br />

Syrop, M. (1975). Leaf curl disease of almond caused by<br />

Taphrina deformans (Berk.) Tul. I. A light microscope<br />

study of the host/parasite relationship. Pro<strong>to</strong>plasma,<br />

85, 39 56.<br />

Syrop, M. J. & Beckett, A. (1972). The origin of ascosporedelimiting<br />

membranes in Taphrina deformans. Archiv<br />

für Mikrobiologie, 86, 185 191.<br />

Syrop, M. J. & Beckett, A. (1976). Leaf curl disease of<br />

almond caused by Taphrina deformans (Berk.) Tul. 3.<br />

Ultrastructural cy<strong>to</strong>logy of the pathogen. Canadian<br />

Journal of Botany, 54, 293 305.<br />

Szabo, S. P., O’Day, D. H. & Chagla, A. H. (1982). Cell<br />

fusion, nuclear fusion, and zygote differentiation<br />

during sexual development of Dictyostelium discoideum.<br />

Developmental Biology, 90, 375 382.<br />

Sziráki, I., Balázs, E. & Király, Z. (1975). Increased levels<br />

of cy<strong>to</strong>kinin and indole-acetic acid in peach leaves<br />

infected with Taphrina deformans. Physiological Plant<br />

Pathology, 5, 45 50.<br />

Takamatsu, S., Hirata, T. & Sa<strong>to</strong>, Y. (2000). A parasitic<br />

transition from trees <strong>to</strong> herbs occurred at least twice<br />

in tribe Cys<strong>to</strong>theceae (Erysiphaceae): evidence from<br />

nuclear ribosomal DNA. Mycological Research, 104,<br />

1304 1311.<br />

Takamatsu, S., Niinomi, S., Cabrera de Álvarez, M. G., et<br />

al. (2005). Caespithotheca gen. nov., an ancestral genus<br />

in the Erysiphales. Mycological Research, 109, 903 911.<br />

Takeshige, K., Baba, M., Tsuboi, S., Noda, T. & Ohsumi,<br />

Y. (1992). Au<strong>to</strong>phagy in yeast demonstrated with<br />

proteinase-deficient mutants and conditions for its<br />

induction. Journal of Cell Biology, 119, 301 311.<br />

Talbot, P. H. B. (1977). The Sirex Amylostereum Pinus<br />

association. Annual Review of Phy<strong>to</strong>pathology, 15,41 54.<br />

Tanabe, Y., O’Donnell, K., Saikawa, M. & Sugiyama, J.<br />

(2000). Molecular phylogeny of parasitic Zygomycota<br />

(Dimargaritales, Zoopagales) based on nuclear small<br />

subunit ribosomal DNA sequences. Molecular<br />

Phylogenetics and Evolution, 16, 253 262.<br />

Tanabe, Y., Saikawa, M., Watanabe, M. M. & Sugiyama,<br />

J. (2004). Molecular phylogeny of Zygomycota based<br />

on EF-1a and RPB1 sequences: limitations and utility<br />

of alternative markers <strong>to</strong> rDNA. Molecular<br />

Phylogenetics and Evolution, 30, 438 449.<br />

Tanabe, Y., Watanabe, M. M. & Sugiyama, J. (2005).<br />

Evolutionary relationships among basal fungi<br />

(Chytridiomycota and Zygomycota): insights from<br />

molecular phylogenetics. Journal of General and<br />

Applied Microbiology, 51, 267 276.<br />

Tanaka, K. (1970). Mi<strong>to</strong>sis in the fungus Basidiobolus<br />

ranarum as revealed by electron microscopy.<br />

Pro<strong>to</strong>plasma, 70, 423 440.<br />

Tanaka, K. & Hirata, A. (1982). Ascospore development<br />

in the fission yeasts Schizosaccharomyces pombe and S.<br />

japonicus. Journal of Cell Science, 56, 263 279.<br />

Tanaka, K. & Kanbe, T. (1986). Mi<strong>to</strong>sis in the fission<br />

yeast Schizosaccharomyces pombe as revealed by freezesubstitution<br />

electron-microscopy. Journal of Cell<br />

Science, 80, 253 268.<br />

Tang, W., Yang, H. & Ryder, M. (2001). Research and<br />

application of Trichoderma spp. in biological control<br />

of plant pathogens. In Bio-Exploitation of Filamen<strong>to</strong>us<br />

<strong>Fungi</strong>, ed. S. B. Pointing & K. B. Hyde. Hong Kong:<br />

Fungal Diversity Press, pp. 403 35.<br />

Tange, Y. & Niwa, O. (1995). A selection system for<br />

diploid and against haploid cells in<br />

Schizosaccharomyces pombe. Molecular and General<br />

Genetics, 248, 644 648.<br />

Tapper, R. (1981). Direct measurement of translocation<br />

of carbohydrate in the lichen, Cladonia convoluta, by<br />

quantitative au<strong>to</strong>radiography. New Phy<strong>to</strong>logist, 89,<br />

429 437.<br />

Tate, K. G. & Wood, P. N. (2000). Potential ascospore<br />

production and resulting blossom blight by Monilinia<br />

fructicola in unsprayed peach trees. New Zealand<br />

Journal of Crop and Horticultural Science, 28, 219 224.<br />

Taylor, B. N., Harrer, T., Pscheidl, E., et al. (2003).<br />

Surveillance of nosocomial transmission of Candida<br />

albicans in an intensive care unit by DNA fingerprinting.<br />

Journal of Hospital Infection, 55, 283 289.<br />

Taylor, D. L. & Bruns, T. D. (1997). Independent specialized<br />

invasions of ec<strong>to</strong>mycorrhizal mutualism by<br />

two nonpho<strong>to</strong>synthetic orchids. Proceedings of the<br />

National Academy of Sciences of the USA, 94, 4510 4515.


802 REFERENCES<br />

Taylor, J. & Birdwell, D. O. (2000). A scanning electron<br />

microscopic study of the infection of water oak<br />

(Quercus nigra) byTaphrina caerulescens. Mycologia, 92,<br />

309 311.<br />

Taylor, T. N., Remy, W. & Hass, H. (1992). <strong>Fungi</strong> from the<br />

Lower Devonian Rhynie chert: Chytridiomycetes.<br />

American Journal of Botany, 79, 1233 1241.<br />

Taylor, T. N., Remy, W., Hass, H. & Kerp, H. (1995). Fossil<br />

arbuscular mycorrhizae from the Early Devonian.<br />

Mycologia, 87, 560 573.<br />

Taylor, T. N., Hass, H. & Kerp, H. (1997). A cyanolichen<br />

from the Lower Devonian Rhynie chert. American<br />

Journal of Botany, 84, 992 1004.<br />

Taylor, T. N., Hass, H. & Kerp, H. (1999). The oldest fossil<br />

ascomycetes. Nature, 399, 648.<br />

Taylor, T. N., Hass, H., Kerp, H., Krings, M. & Hanlin,<br />

R. T. (2005). Perithecial ascomycetes from the 400<br />

million year old Rhynie chert: an example of<br />

ancestral polymorphism. Mycologia, 97, 269 285.<br />

Tehler, A. (1996). Systematics, phylogeny and classification.<br />

In Lichen Biology, ed. T. H. Nash. Cambridge:<br />

Cambridge University Press, pp. 217 239.<br />

Tehler, A., Little, D. P. & Farris, J. S. (2003). The fulllength<br />

phylogenetic tree from 1551 ribosomal<br />

sequences of chitinous fungi. <strong>Fungi</strong>. Mycological<br />

Research, 107, 901 916.<br />

Teichert, U., Mechler, B., Müller, H. & Wolf, D. H. (1989).<br />

Lysosomal (vacuolar) proteinases of yeast are essential<br />

catalysts for protein degradation, differentiation,<br />

and cell survival. Journal of Biological Chemistry,<br />

264, 16 037 16 045.<br />

Temmink, J. H. M. & Campbell, R. N. (1968). The ultrastructure<br />

of Olpidium brassicae. I. Formation of<br />

sporangia. Canadian Journal of Botany, 46, 951 955.<br />

Temmink, J. H. M. & Campbell, R. N. (1969a). The<br />

ultrastructure of Olpidium brassicae. II. Zoospores.<br />

Canadian Journal of Botany, 47, 227 231.<br />

Temmink, J. H. M. & Campbell, R. N. (1969b). The<br />

ultrastructure of Olpidium brassicae. III. Infection of<br />

host roots. Canadian Journal of Botany, 47, 421 424.<br />

Temmink, J. H. M., Campbell, R. N. & Smith, P. R. (1970).<br />

Specificity and site of in vitro acquisition of <strong>to</strong>bacco<br />

necrosis virus by zoospores of Olpidium brassicae.<br />

Journal of General Virology, 9, 201 203.<br />

Temple, B., Horgen, P. A., Bernier, L. & Hertz, W. E.<br />

(1997). Cera<strong>to</strong>-ulmin, a hydrophobin secreted by the<br />

causal agents of Dutch elm disease, is a parasitic<br />

fitness fac<strong>to</strong>r. Fungal Genetics and Biology, 22, 39 53.<br />

Tenberge, K. B. (1999). Biology and strategy of the ergot<br />

fungi. In Ergot. The Genus Claviceps, ed. V. Křen & L.<br />

Cvak. Amsterdam: Harwood Academic Publishers,<br />

pp. 25 56.<br />

ten Have, R. & Teunissen, P. J. M. (2001). Oxidative<br />

mechanisms involved in lignin<br />

degradation by white-rot fungi. Chemical Reviews,<br />

101, 3397 3413.<br />

Tenney, K., Hunt, I., Sweigard, J., et al. (2000). hex-1, a<br />

gene unique <strong>to</strong> filamen<strong>to</strong>us fungi, encodes the<br />

major protein of the Woronin body and functions as<br />

a plug for septal pores. Fungal Genetics and Biology, 31,<br />

205 217.<br />

Tereshina, V. M. & Feofilova, E. P. (1995). Two sporulation<br />

types in the Mucorous fungus Blakeslea trispora<br />

morphological characteristics. Microbiology, 64,<br />

580 583.<br />

Terhune, B. T. & Hoch, H. C. (1993). Substrate hydrophobicity<br />

and adhesion of Uromyces urediospores and<br />

germlings. Experimental Mycology, 17, 241 252.<br />

Terhune, B. T., Allen, E. A., Hoch, H. C., Wergin, W. P. &<br />

Erbe, E. F. (1991). S<strong>to</strong>matal on<strong>to</strong>geny and morphology<br />

in Phaseolus vulgaris in relation <strong>to</strong> infection<br />

structure initiation in Uromyces appendiculatus.<br />

Canadian Journal of Botany, 69, 477 484.<br />

Teunissen, H. A. S., Verkooijen, J., Cornelissen, B. J. C. &<br />

Haring, M. A. (2002). Genetic exchange of avirulence<br />

determinants and extensive karyotype rearrangements<br />

in parasexual recombinants of Fusarium<br />

oxysporum. Molecular Genetics and Genomics, 268,<br />

298 310.<br />

Thaxter, R. (1888). The En<strong>to</strong>mophthoreae of the United<br />

States. Memoirs of the Bos<strong>to</strong>n Society for Natural His<strong>to</strong>ry,<br />

4, 133 201.<br />

Theodorou, M. K., Lowe, S. E. & Trinci, A. P. J. (1992).<br />

Anaerobic fungi and the rumen ecosystem. In The<br />

Fungal Community. Its Organization and Role in the<br />

Ecosystem, 2nd edn, ed. G. C. Carroll & D. T. Wicklow.<br />

New York: Marcel Dekker Inc., pp. 43 72.<br />

Theodorou, M. K., Zhu, W.-Y., Rickers, A., et al. (1996).<br />

Biochemistry and ecology of anaerobic fungi. In The<br />

Mycota VI: Human and Animal Relationships, ed. D. H.<br />

Howard & J. D. Miller. Berlin: Springer-Verlag,<br />

pp. 265 295.<br />

Thielke, C. (1982). Meiotic divisions in the basidium. In<br />

Basidium and Basidiocarp: Evolution, Cy<strong>to</strong>logy, Function<br />

and Development, ed. K. Wells & E. K. Wells. New York:<br />

Springer-Verlag, pp. 75 91.<br />

Thiers, H. D. (1984). The secotioid syndrome. Mycologia,<br />

76, 1 8.<br />

Thines, E., Eilbert, F., Sterner, O. & Anke, H. (1997).<br />

Signal transduction leading <strong>to</strong> appressorium formation<br />

in germinating conidia of Magnaporthe grisea:<br />

effects of second messengers diacylglycerols, ceramides<br />

and sphingomyelin. FEMS Microbiology Letters,<br />

156, 91 94.


REFERENCES<br />

803<br />

Thines, E., Weber, R. W. S. & Talbot, N. J. (2000). MAP<br />

kinase and protein kinase A-dependent mobilization<br />

of triacylglycerol and glycogen during appressorium<br />

turgor generation by Magnaporthe grisea. Plant Cell,<br />

12, 1703 1718.<br />

Thirumalachar, M. J. & Dickson, J. G. (1949).<br />

Chlamydospore germination, nuclear cycle and<br />

artificial culture of Urocystis agropyri on red <strong>to</strong>p.<br />

Phy<strong>to</strong>pathology, 39, 333 339.<br />

Thomas, D. des S. & McMorris, T. C. (1987). Allomonal<br />

functions of the steroid hormone, antheridiol, in<br />

water mold Achlya. Journal of Chemical Ecology, 13,<br />

1131 1137.<br />

Thomas, P. L. (1984). Recombination of virulence genes<br />

following hybridization between isolates of Ustilago<br />

nuda infecting barley under natural conditions.<br />

Canadian Journal of Plant Pathology, 6, 101 104.<br />

Thomas, P. L. & Menzies, J. G. (1997). Cereal smuts in<br />

Mani<strong>to</strong>ba and Saskatchewan, 1989 1995. Canadian<br />

Journal of Plant Pathology, 19, 161 165.<br />

Thomma, B. P. H. J. (2003). Alternaria spp.: from general<br />

saprophyte <strong>to</strong> specific parasite. Molecular Plant<br />

Pathology, 4, 225 236.<br />

Thompson, J. E., Fahnes<strong>to</strong>ck, S., Farrall, L., et al. (2000).<br />

The second naphthol reductase of fungal melanin<br />

biosynthesis in Magnaporthe grisea. Journal of Biological<br />

Chemistry, 275, 34 867 34 872.<br />

Thompson, W. & Rayner, A. D. M. (1982). Spatial<br />

structure in a population of Tricholomopsis platyphylla<br />

in a woodland site. New Phy<strong>to</strong>logist, 92, 103 114.<br />

Thompson-Coffe, C. & Zickler, D. (1994). How the<br />

cy<strong>to</strong>skele<strong>to</strong>n recognizes and sorts nuclei of opposite<br />

mating-type during the sexual cycle in filamen<strong>to</strong>us<br />

ascomycetes. Developmental Biology, 165, 257 271.<br />

Thomson, J., Melville, L. H. & Peterson, R. L. (1989).<br />

Interaction between the ec<strong>to</strong>mycorrhizal fungus<br />

Pisolithus tinc<strong>to</strong>rius and root hairs of Picea mariana (P).<br />

American Journal of Botany, 76, 632 636.<br />

Thorn, R. G., Moncalvo, J. M., Reddy, C. A. & Vilgalys, R.<br />

(2000). Phylogenetic analyses and the distribution of<br />

nema<strong>to</strong>phagy support a monophyletic Pleurotaceae<br />

within the polyphyletic pleuro<strong>to</strong>id lentinoid fungi.<br />

Mycologia, 92, 241 252.<br />

Thumm, M. (2000). Structure and function of the yeast<br />

vacuole and its role in au<strong>to</strong>phagy. Microscopy Research<br />

and Technique, 51, 563 572.<br />

Tils<strong>to</strong>n, E. L., Pitt, D. & Groenhof, A. C. (2002).<br />

Composted recycled organic matter suppresses soilborne<br />

diseases of field crops. New Phy<strong>to</strong>logist, 154,<br />

731 740.<br />

Todd, N. K. & Rayner, A. D. M. (1980). Fungal individualism.<br />

Science Progress, 66, 331 354.<br />

Tomiyama, S., Sakuma, T., Ishizaka, N., et al. (1968).<br />

A new antifungal substance isolated from resistant<br />

pota<strong>to</strong> tuber tissue infected by pathogens.<br />

Phy<strong>to</strong>pathology, 58, 115 116.<br />

Tomlinson, J. A. (1958). Crook root of watercress. II. The<br />

control of the disease with zinc-fritted glass and the<br />

mechanism of its action. Annals of Applied Biology, 46,<br />

608 621.<br />

Tommerup, I. C. & Broadbent, D. (1974). Nuclear fusion,<br />

meiosis and the origin of dikaryotic hyphae in<br />

Armillaria mellea. Archiv für Mikrobiologie, 103,<br />

279 282.<br />

Tommerup, I. C. & Ingram, D. S. (1971). The life cycle of<br />

Plasmodiophora brassicae Woron. in Brassica tissue<br />

cultures and in intact roots. New Phy<strong>to</strong>logist, 70,<br />

327 332.<br />

Tommerup, I. C. & Sivasithamparam, K. (1990).<br />

Zygospores and asexual spores of Gigaspora decipiens,<br />

an arbuscular mycorrhizal fungus. Mycological<br />

Research, 94, 897 900.<br />

Ton, J., van Pelt, J. A., van Loon, L. C. & Pieterse, C. M.<br />

(2002). Differential effectiveness of salicylate-dependent<br />

and jasmonate/ethylene-dependent induced<br />

resistance in Arabidopsis. Molecular Plant Microbe<br />

Interactions, 15, 27 34.<br />

Torralba, S., Raudaskoski, M. & Pedregosa, A. M. (1998).<br />

Effects of methyl benzimidazole-2-yl carbamate on<br />

microtubule and actin cy<strong>to</strong>skele<strong>to</strong>n in Aspergillus<br />

nidulans. Pro<strong>to</strong>plasma, 202, 54 64.<br />

Torralba, S., Heath, I. B. & Ottensmeyer, F. P. (2001).<br />

Ca 2þ shuttling in vesicles during tip growth in<br />

Neurospora crassa. Fungal Genetics and Biology, 33,<br />

181 193.<br />

Torralba, S., Pisabarro, A. G. & Ramirez, L. (2004).<br />

Immunofluorescence microscopy of the microtubule<br />

cy<strong>to</strong>skele<strong>to</strong>n during conjugate division in the<br />

dikaryon of Pleurotus ostreatus N001. Mycologia, 96,<br />

41 51.<br />

Tournas, V. (1994). Heat-resistant fungi of importance<br />

<strong>to</strong> the food and beverage industry. Critical Reviews in<br />

Microbiology, 20, 243 263.<br />

Townsend, B. B. (1954). Morphology and development<br />

of fungal rhizomorphs. Transactions of the British<br />

Mycological Society, 37, 222 233.<br />

Townsend, B. B. & Willetts, H. J. (1954). The development<br />

of sclerotia of certain fungi. Transactions of the<br />

British Mycological Society, 37, 213 221.<br />

Trail, F., Gaffoor, I. & Vogel, S. (2005).<br />

Ejection mechanisms and trajec<strong>to</strong>ry of the<br />

ascospores of Gibberella zeae (anamorph Fusarium<br />

graminearum). Fungal Genetics and Biology, 42,<br />

528 533.


804 REFERENCES<br />

Trappe, J. M. (1979). The orders, families, and genera of<br />

hypogeous Ascomycotina (truffles and their relatives).<br />

Mycotaxon, 9, 297 340.<br />

Trappe, J. M. & Maser, C. (1977). Ec<strong>to</strong>mycorrhizal fungi:<br />

interactions of mushrooms and truffles with beasts<br />

and trees. In Mushrooms and Man, an Interdisciplinary<br />

Approach <strong>to</strong> Mycology, ed. T. Walters. Albany, Oregon:<br />

USDA Forestry Service, pp. 165 179.<br />

Tredway, L. P., White, J. F., Gaut, B. S., et al. (1999).<br />

Phylogenetic relationships within and between<br />

Epichloe and Neotyphodium endophytes as estimated<br />

by AFLP markers and rDNA sequences. Mycological<br />

Research, 103, 1593 1603.<br />

Trembley, M. L., Ringli, C. & Honegger, R. (2002).<br />

Morphological and molecular analysis of early stages<br />

in the resynthesis of the lichen Baeomyces rufus.<br />

Mycological Research, 106, 768 776.<br />

Tribe, H. T. (1955). Studies on the physiology of<br />

parasitism. XIX. On the killing of plant cells by<br />

enzymes from Botrytis cinerea and Bacterium aroideae.<br />

Annals of Botany N.S., 19, 351 368.<br />

Tribe, H. T. (1957). On the parasitism of Sclerotinia<br />

trifoliorum by Coniothyrium minitans. Transactions of the<br />

British Mycological Society, 40, 489 499.<br />

Tribe, H. T. (1977). On ‘Olpidium nema<strong>to</strong>deae Skvortzow’.<br />

Transactions of the British Mycological Society, 69,<br />

509 511.<br />

Tribe, H. T. & Weber, R. W. S. (2002). A low-temperature<br />

fungus from cardboard, Myxotrichum chartarum.<br />

Mycologist, 16, 3 5.<br />

Trinci, A. P. J. (1991). ‘Quorn’ mycoprotein. Mycologist, 5,<br />

106 109.<br />

Trinci, A. P. J., Davies, D. R., Gull, K., et al. (1994).<br />

Anaerobic fungi in herbivorous animals. Mycological<br />

Research, 98, 129 152.<br />

Trione, E. J. (1972). Isolation of Tilletia caries from<br />

infected wheat plants. Phy<strong>to</strong>pathology, 62, 1096 1097.<br />

Tsai, H.-F., Liu, J.-S., Staben, C., et al. (1994). Evolutionary<br />

classification of fungal endophytes of tall fescue<br />

grass by hydridization of Epichloe species. Proceedings<br />

of the National Academy of Sciences of the USA, 91,<br />

2542 2546.<br />

Tsao, P. H. (1983). Fac<strong>to</strong>rs affecting isolation and<br />

quantitation of Phy<strong>to</strong>phthora from soil. In<br />

Phy<strong>to</strong>phthora. Its Biology, Taxonomy, Ecology, and<br />

Pathology, ed. D. C. Erwin, S. Bartnicki-Garcia &<br />

P. H. Tsao. St. Paul: APS Press, pp. 219 236.<br />

Tsuneda, A. & Currah, R. S. (2004). Ascomatal<br />

morphogenesis in Myxotrichum arcticum<br />

supports the derivation of the Myxotrichaceae<br />

from a discomyce<strong>to</strong>us ances<strong>to</strong>r. Mycologia, 96,<br />

627 635.<br />

Tu, C. C., Kimbrough, J. W. & Aldrich, H. C. (1977).<br />

Cy<strong>to</strong>logy and ultrastructure of Thanatephorus and<br />

related taxa of the Rhizoc<strong>to</strong>nia complex. Canadian<br />

Journal of Botany, 55, 2419 2436.<br />

Tu, J. C. (1985). Tolerance of white bean (Phaseolus<br />

vulgaris) <strong>to</strong> white mold (Sclerotinia sclerotiorum) associated<br />

with <strong>to</strong>lerance <strong>to</strong> oxalic acid. Physiological Plant<br />

Pathology, 26, 111 117.<br />

Tu, J. C. (1988). The role of white mold-infected white<br />

bean (Phaseolus vulgaris L.) seeds in the dissemination<br />

of Sclerotinia sclerotiorum (Lib.) de Bary.<br />

Phy<strong>to</strong>pathologische Zeitschrift, 121, 40 50.<br />

Tucker, S. L. & Talbot, N. J. (2001). Surface attachment<br />

and pre-penetration stage development by plant<br />

pathogenic fungi. Annual Review of Phy<strong>to</strong>pathology, 39,<br />

385 417.<br />

Tudzynski, B. (1997). Fungal phy<strong>to</strong>hormones in pathogenic<br />

and mutualistic associations. In The Mycota VA:<br />

Plant Relationships, ed. G. C. Carroll & P. Tudzynski.<br />

Berlin: Springer-Verlag, pp. 167 184.<br />

Tudzynski, P. (1999). Genetics of Claviceps purpurea. In<br />

Ergot. The Genus Claviceps, ed. V. Křen & L. Cvak.<br />

Amsterdam: Harwood Academic Publishers,<br />

pp. 79 93.<br />

Tudzynski, P., Correia, T. & Keller, U. (2001).<br />

Biotechnology and genetics of ergot<br />

alkaloids. Applied Microbiology and Biotechnology,<br />

57, 593 605.<br />

Tulasne, L. R. & Tulasne, C. C. (1931). Selecta Fungorum<br />

Carpologia I (English translation by W. B. Grove).<br />

Oxford: Clarendon Press.<br />

Tunlid, A., Jansson, H.-B. & Nordbring-Hertz, B. (1992).<br />

Fungal attachment <strong>to</strong> nema<strong>to</strong>des. Mycological<br />

Research, 96, 401 412.<br />

Tuno, N. (1998). Spore dispersal of Dictyophora fungi<br />

(Phallaceae) by flies. Ecological Research, 13, 7 15.<br />

Tuno, N. (1999). Insect feeding on spores of a bracket<br />

fungus, Elfvingia applanata (Pers.) Karst.<br />

(Ganodermataceae, Aphyllophorales). Ecological<br />

Research, 14, 97 103.<br />

Turner, B. C., Perkins, D. D. & Fairfield, A. (2001).<br />

Neurospora from natural populations: a global study.<br />

Fungal Genetics and Biology, 32, 67 92.<br />

Tuse, D. (1984). Single-cell protein: current status and<br />

future prospects. Critical Reviews in Food Science and<br />

Nutrition, 19, 273 325.<br />

Tyler, B. M. (2002). Molecular basis of recognition<br />

between Phy<strong>to</strong>phthora pathogens and their hosts.<br />

Annual Review of Phy<strong>to</strong>pathology, 40, 137 167.<br />

Tyrell, D. & MacLeod, D. M. (1975). In vitro germination<br />

of En<strong>to</strong>mophthora aphidis resting spores. Canadian<br />

Journal of Botany, 53, 1188 1191.


REFERENCES<br />

805<br />

Uchida, W., Matsunaga, S., Sugiyama, R., Kazama, Y. &<br />

Kawano, S. (2003). Morphological development of<br />

anthers induced by the dimorphic smut fungus<br />

Microbotryum violaceum in female flowers of<br />

the dioecious plant Silene latifolia. Planta, 218,<br />

240 248.<br />

Uecker, F. A. (1976). Development and cy<strong>to</strong>logy of<br />

Sordaria humana. Mycologia, 68, 30 46.<br />

Uecker, F. A. (1988). A World List of Phomopsis Names with<br />

Notes on Nomenclature, Morphology and Biology. Berlin:<br />

J. Cramer.<br />

Ueng, P. P., Subramaniam, K., Chen, W., et al. (1998).<br />

Intraspecific genetic variation of Stagonospora avenae<br />

and its differentiation from S. nodorum. Mycological<br />

Research, 102, 607 614.<br />

Uesugi, Y. (1998). <strong>Fungi</strong>cide classes: chemistry, uses and<br />

mode of action. In <strong>Fungi</strong>cidal Activity. Chemical and<br />

Biological Approaches <strong>to</strong> Plant Protection, ed. D. Hutson &<br />

J. Miyamo<strong>to</strong>. Chichester: John Wiley & Sons,<br />

pp. 23 56.<br />

Uetake, Y., Kobayashi, K. & Ogoshi, A. (1992).<br />

Ultrastructural changes during the symbiotic development<br />

of Spiranthes sinensis (Orchidaceae) pro<strong>to</strong>corms<br />

associated with binucleate Rhizoc<strong>to</strong>nia<br />

anas<strong>to</strong>mosis group C. Mycological Research, 96,<br />

199 209.<br />

Ugadawa, S. (1984). Taxonomy of myco<strong>to</strong>xin-producing<br />

Chae<strong>to</strong>mium. In Toxigenic <strong>Fungi</strong>. Their Toxins and Health<br />

Hazards, ed. H. Kura<strong>to</strong> & Y. Ueno. Tokyo: Elsevier,<br />

pp. 139 47.<br />

Uhm, J. Y. & Fujii, H. (1983a). Ascospore dimorphism in<br />

Sclerotinia trifoliorum and cultural characters of<br />

strains from different-sized spores. Phy<strong>to</strong>pathology,<br />

73, 565 569.<br />

Uhm, J. Y. & Fujii, H. (1983b). Heterothallism and<br />

mating type mutation in Sclerotinia trifoliorum.<br />

Phy<strong>to</strong>pathology, 73, 569 572.<br />

Ulken, A. & Sparrow, F. K. (1968). Estimation of chytrid<br />

propagules in Douglas Lake by the MPN-Pollen grain<br />

method. Veröffentlichungen des Instituts für<br />

Meeresforschung Bremerhaven, 11, 83 88.<br />

Ullrich, R. C. & Anderson, J. B. (1978). Sex and diploidy<br />

in Armillaria mellea. Experimental Mycology, 2,<br />

119 129.<br />

Ullrich, R. C., Specht, C. A., Stankis, M. M., Giasson, L. &<br />

Novotny, C. P. (1991). Molecular biology of matingtype<br />

determination in Schizophyllum commune. In<br />

Genetic Engineering, Principles and Methods 13, ed.<br />

J. K. Setlow. New York: Plenum Press, pp. 279 306.<br />

Ullstrup, A. J. (1972). The impacts of the southern corn<br />

leaf blight epidemics of 1970 71. Annual Review of<br />

Phy<strong>to</strong>pathology, 10, 37 50.<br />

Umar, M. H. & van Griensven, L. J. L. D. (1997).<br />

Morphogenetic cell death in developing primordia<br />

of Agaricus bisporus. Mycologia, 89, 274 277.<br />

Untereiner, W. A. (2000). Capronia and its anamorphs:<br />

exploring the value of morphological and molecular<br />

characters in the systematics of the<br />

Herpotrichiellaceae. Studies in Mycology, 45, 141 149.<br />

Untereiner, W. A., Scott, J. A., Naveau, F. A., et al. (2004).<br />

The Ajellomycetaceae, a new family of vertebrateassociated<br />

Onygenales. Mycologia, 96, 812 821.<br />

Urashima, A. S., Igirashi, S. & Ka<strong>to</strong>, H. (1993). Host<br />

range, mating type, and fertility of Pyricularia grisea<br />

from wheat in Brazil. Plant Disease, 77, 1211 1261.<br />

Urban, A., Neuner-Plattner, I., Krisai-Greilhuber, I. &<br />

Haselwandter, K. (2004). Molecular studies on terricolous<br />

microfungi reveal novel anamorphs of two<br />

Tuber species. Mycological Research, 108, 749 758.<br />

Urban, M., Bhargava, T. & Hamer, J. E. (1999). An ATPdriven<br />

efflux pump is a novel pathogenicity fac<strong>to</strong>r in<br />

rice blast disease. EMBO Journal, 18, 512 521.<br />

Urbanus, J. F., van den Ende, H. & Koch, B. (1978).<br />

Calcium oxalate crystals in the wall of Mucor mucedo.<br />

Mycologia, 70, 829 842.<br />

Urquhart, E. J. & Punja, Z. K. (2002). Hydrolytic enzymes<br />

and antifungal compounds produced by Tilletiopsis<br />

species, phyllosphere yeasts that are antagonists of<br />

powdery mildews. Canadian Journal of Microbiology, 48,<br />

219 229.<br />

Valent, B. (1997). The rice blast fungus, Magnaporthe<br />

grisea. InThe Mycota VB: Plant Relationships, ed. G. C.<br />

Carroll & P. Tudzynski. Berlin: Springer-Verlag,<br />

pp. 37 54.<br />

van Brummelen, J. (1967). A world monograph of the<br />

genera Ascobolus and Saccobolus (Ascomycetes,<br />

Pezizales). Persoonia Supplement 1.<br />

van Brummelen, J. (1981). The operculate ascus and<br />

allied forms. In Ascomycete Systematics. The Luttrellian<br />

Concept, ed. D. R. Reynolds. New York: Springer-<br />

Verlag, pp. 27 48.<br />

van den Bogart, H. G., van den Ende, G., van Loon, P. C.<br />

& van Griensven, L. J. (1993). Mushroom worker’s<br />

lung: serologic reactions <strong>to</strong> thermophilic actinomycetes<br />

present in the air of compost tunnels.<br />

Mycopathologia, 122, 21 28.<br />

van der Auwera, G., de Baere, R., van de Peer, Y., et al.<br />

(1995). The phylogeny of the Hyphochytriomycota as<br />

deduced from ribosomal RNA sequences of<br />

Hyphochytrium catenoides. Molecular Biology and<br />

Evolution, 12, 671 678.<br />

Vanderhaegen, B., Neven, H., Coghe, S., et al. (2003).<br />

Bioflavoring and beer refermentation. Applied<br />

Microbiology and Biotechnology, 62, 140 150.


806 REFERENCES<br />

van der Plaats-Niterink, A. J. (1981). Monograph of the<br />

genus Pythium. Studies in Mycology, 21, 1 242. Baarn:<br />

Centraalbureau voor Schimmelcultures.<br />

van der Plank, J. E. (1963). Plant Diseases: Epidemics and<br />

Control. New York: Academic Press.<br />

van der Walt, J. P. (1970). The perfect and imperfect<br />

states of Sporobolomyces salmonicolor. An<strong>to</strong>nie van<br />

Leeuwenhoek, 36, 49 55.<br />

van Donk, E. & Bruning, K. (1992). Ecology of Aquatic<br />

<strong>Fungi</strong> In and On Algae. Bris<strong>to</strong>l: Biopress Ltd.<br />

van Eijik, G. W. & Roeymans, H. J. (1982). Distribution of<br />

carotenoids and sterols in relation <strong>to</strong> the taxonomy<br />

of Taphrina and Pro<strong>to</strong>myces. An<strong>to</strong>nie van Leeuwenhoek,<br />

48, 257 264.<br />

van Griensven, L. J. L. D., ed. (1988). The Cultivation of<br />

Mushrooms. Rusting<strong>to</strong>n, UK: Darling<strong>to</strong>n Mushroom<br />

Labora<strong>to</strong>ries Ltd.<br />

Vanittanakom, N., Cooper, C. R., Fisher, M. C. &<br />

Sirisanthana, T. (2006). Penicillium marneffei infection<br />

and recent advances in the epidemiology and<br />

molecular biology aspects. Clinical Microbiology<br />

Reviews, 19, 95 110.<br />

Vánky, K. (1987). Illustrated Genera of Smut <strong>Fungi</strong>.<br />

Stuttgart: Gustav Fischer.<br />

Vánky, K. (1994). European Smut <strong>Fungi</strong>. Stuttgart: Gustav<br />

Fischer.<br />

Vánky, K. (1998). The genus Microbotryum (smut fungi).<br />

Mycotaxon, 67, 33 60.<br />

Vánky, K. (1999). The new classifica<strong>to</strong>ry system for<br />

smut fungi, and two new genera. Mycotaxon, 70,<br />

35 49.<br />

van Leeuwen, G. C. M., Baayen, R. P., Holb, I. J. & Jeger,<br />

M. J. (2002). Distinction of the Asiatic brown rot<br />

fungus Monilia polystroma sp. nov. from M. fructigena.<br />

Mycological Research, 106, 444 451.<br />

van Oorschot, C. A. N. (1985). Taxonomy of the<br />

Dactylaria complex. V. A review of Arthrobotrys and<br />

allied genera. Studies in Mycology, 26, 61 96.<br />

van Rinsum, J., Klis, F. M. & van den Ende, H. (1991). Cell<br />

wall glucomannoproteins of Saccharomyces cerevisiae<br />

mnn9. Yeast, 7, 717 726.<br />

Vanterpool, C. T. (1959). Oospore germination in Albugo<br />

candida. Canadian Journal of Botany, 37, 169 172.<br />

van Vuuren, H. J. J. & Jacobs, C. J. (1992). Killer yeasts in<br />

the wine industry: a review. American Journal of<br />

Enology and Viticulture, 43, 119 128.<br />

Varga, J., Rigo, K., Toth, B., Teren, J. & Kozakiewicz, Z.<br />

(2003). Evolutionary relationships among Aspergillus<br />

species producing economically important myco<strong>to</strong>xins.<br />

Food Technology and Biotechnology, 41, 29 36.<br />

Varitchak, B. (1931). Contribution à l’étude du développement<br />

des Ascomycètes. Botaniste, 23, 1 123.<br />

Vasiliauskas, R., Lygis, V., Thor, M. & Stenlid, J. (2004).<br />

Impact of biological (Rots<strong>to</strong>p) and chemical (urea)<br />

treatments on fungal community structure in<br />

freshly cut Picea abies stumps. Biological Control, 31,<br />

405 413.<br />

Vasiliauskas, R., Larsson, K. H., Larsson, K. H. &<br />

Stenlid, J. (2005). Persistence and long-term impact<br />

of Rots<strong>to</strong>p biological control agent on<br />

mycodiversity in Picea abies stumps. Biological Control,<br />

32, 295 304.<br />

Vaughan-Martini, A. & Martini, A. (1998a).<br />

Schizosaccharomyces Lindner. In The Yeasts. A Taxonomic<br />

Study, 4th edn, ed. C. P. Kurtzman & J. W. Fell.<br />

Amsterdam: Elsevier, pp. 391 4.<br />

Vaughan-Martini, A. & Martini, A. (1998b).<br />

Saccharomyces Meyen ex Reess. In The Yeasts: A<br />

Taxonomic Study, 4th edn, ed. C. P. Kurtzman & J. W.<br />

Fell. Amsterdam: Elsevier, pp. 358 71.<br />

Veenhuis, M., Nordbring-Hertz, B. & Harder, W. (1985).<br />

An electron microscopical analysis of capture and<br />

initial stages of penetration of nema<strong>to</strong>des by<br />

Arthrobotrys oligospora. An<strong>to</strong>nie van Leeuwenhoek, 51,<br />

385 398.<br />

Veenhuis, M., van Wijk, C., Wyss, U., Nordbring-Hertz,<br />

B. & Harder, W. (1989). Significance of electron dense<br />

microbodies in trap cells of the nema<strong>to</strong>phagous<br />

fungus Arthrobotrys oligospora. An<strong>to</strong>nie van<br />

Leeuwenhoek, 56, 251 261.<br />

Vellinga, E. C. (2004). Genera in the family Agaricaceae:<br />

evidence from nrITS and nrLSU sequences.<br />

Mycological Research, 108, 354 377.<br />

Vellinga, E. C., De Kok, R. P. J. & Bruns, T. D. (2003).<br />

Phylogeny and taxonomy of Macrolepiota<br />

(Agaricaceae). Mycologia, 95, 442 456.<br />

Verkley, G. J. M. (1992). Ultrastructure of the apical<br />

apparatus of asci in Ombrophila violacea, Neobulgaria<br />

pura and Bulgaria inquinans (Leotiales). Persoonia, 15,<br />

3 22.<br />

Verkley, G. J. M. (1993). Ultrastructure of the ascus<br />

apical apparatus in ten species of Sclerotiniaceae.<br />

Mycological Research, 97, 179 194.<br />

Verkley, G. J. M. (1994). Ultrastructure of the ascus<br />

apical apparatus in Leotia lubrica and some<br />

Geoglossaceae (Leotiales, Ascomycotina). Persoonia,<br />

15, 405 430.<br />

Verkley, G. J. M. (1996). Ultrastructure of the ascus in<br />

the genera Lachnum and Trichopeziza<br />

(Hyaloscyphaceae, Ascomycotina). Nova Hedwigia, 63,<br />

215 228.<br />

Versele, M. & Thorner, J. (2005). Some assembly<br />

required: yeast septins provide the instruction<br />

manual. Trends in Cell Biology, 15, 414 424.


REFERENCES<br />

807<br />

Verstrepen, K. J., Derdelinckx, G., Verachtert, H. &<br />

Delvaux, F. R. (2003). Yeast flocculation: what<br />

brewers should know. Applied Microbiology and<br />

Biotechnology, 61, 197 205.<br />

Viani, R. (2002). Effect of processing on ochra<strong>to</strong>xin A<br />

(OTA) content of coffee. Advances in Experimental<br />

Medicine and Biology, 504, 189 193.<br />

Vida, T. A. & Emr, S. D. (1995). A new vital stain for<br />

visualizing vacuolar membrane dynamics and<br />

endocy<strong>to</strong>sis in yeast. Journal of Cell Biology, 128,<br />

779 792.<br />

Vilgalys, R. & Cubeta, M. A. (1994). Molecular systematics<br />

and population biology of Rhizoc<strong>to</strong>nia. Annual<br />

Review of Phy<strong>to</strong>pathology, 32, 135 155.<br />

Vilgalys, R. & Gonzalez, D. (1990). Organization of<br />

ribosomal DNA in the basidiomycete Thanatephorus<br />

praticola. Current Genetics, 18, 277 280.<br />

Vinogradov, E., Petersen, B. O., Duus, J. Ø. & Wasser, S.<br />

(2004). The structure of the glucuronoxylomannan<br />

produced by culinary medicinal yellow brain<br />

mushroom (Tremella mesenterica Ritz.:Fr.,<br />

Heterobasidiomycetes) grown as one cell biomass in<br />

submerged culture. Carbohydrate Research, 339,<br />

1483 1489.<br />

Vitkov, L., Krautgartner, W. D., Hannig, M., Weitgasser,<br />

R. & S<strong>to</strong>iber, W. (2002). Candida attachment <strong>to</strong> oral<br />

epithelium. Oral Microbiology Immunology, 17, 60 64.<br />

Voegele, R. T. & Mendgen, K. (2003). Rust haus<strong>to</strong>ria:<br />

nutrient uptake and beyond. New Phy<strong>to</strong>logist, 159,<br />

93 100.<br />

Vogel, H. J. (1964). Distribution of lysine pathways<br />

among fungi: evolutionary implications. American<br />

Naturalist, 98, 435 446.<br />

Vogel, S. (2005). Living in a physical world. III. Getting<br />

up <strong>to</strong> speed. Journal of Bioscience, 30, 303 312.<br />

Vogel, J. & Somerville, S. (2002). Powdery mildew of<br />

Arabidopsis: a model system for host parasite interactions.<br />

In The Powdery Mildews. A Comprehensive<br />

Treatise, ed. R. R. Bélanger, W. R. Bushnell, A. J. Dik &<br />

T. L. W. Carver. St. Paul, USA: APS Press, pp. 161 168.<br />

Vogler, D. R. & Bruns, T. D. (1998). Phylogenetic relationships<br />

among the pine stem rust fungi<br />

(Cronartium and Peridermium spp.). Mycologia, 90,<br />

244 257.<br />

Voglmayr, H. (2000). Die aero-aquatischen Pilze des<br />

Sauwaldgebietes. Beiträge zur Naturkunde<br />

Oberösterreichs, 9, 705 728.<br />

Voglmayr, H., Bonner, L. M. & Dick, M. W. (1999).<br />

Taxonomy and oogonial ultrastructure of a new<br />

aero-aquatic peronosporomycete (Medusoides gen.<br />

nov.; Pythioge<strong>to</strong>naceae fam. nov.). Mycological<br />

Research, 103, 591 606.<br />

Volk, T. J. & Leonard, T. J. (1989). Experimental studies<br />

on morel. I. Heterokaryon formation between<br />

monoascosporous strains of Morchella. Mycologia, 81,<br />

523 531.<br />

Volk, T. J. & Leonard, T. J. (1990). Cy<strong>to</strong>logy of the life<br />

cycle of Morchella. Mycological Research, 94, 399 406.<br />

von Arx, J. A. (1950). Über die Ascusform von<br />

Cladosporium herbarum. Sydowia, 4, 320 324.<br />

von Arx, J. A. (1957). Die Arten der Gattung<br />

Colle<strong>to</strong>trichum Cda. Phy<strong>to</strong>pathologische Zeitschrift, 29,<br />

413 468.<br />

von Arx, J. A. (1986). The ascomycete genus Gymnoascus.<br />

Persoonia, 13, 173 183.<br />

von Arx, J. A., Guarro, J. & Figueras, M. J. (1986). The<br />

Ascomycete genus Chae<strong>to</strong>mium. Nova Hedwigia,<br />

Beihefte, 84, 1 117.<br />

von Röpenack, E., Parr, A. & Schulze-Lefert, P. (1998).<br />

Structural analyses and dynamics of soluble and cell<br />

wall-bound phenolics in a broad spectrum resistance<br />

<strong>to</strong> the powdery mildew fungus in barley. Journal of<br />

Biological Chemistry, 273, 9013 9022.<br />

Voos, J. R. (1969). Morphology and life cycle of a new<br />

chytrid with aerial sporangia. American Journal of<br />

Botany, 56, 898 909.<br />

Wagner, T. & Fischer, M. (2001). Natural groups and a<br />

revised system for the European poroid<br />

Hymenochaetales (Basidiomycota) supported by nLSU<br />

rDNA sequence data. Mycological Research, 105,<br />

773 782.<br />

Wahl, I., Anikster, Y., Manisterski, J. & Segal, A. (1984).<br />

Evolution at the center of origin. In The Cereal Rusts 1:<br />

Origins, Specificity, Structure, and Physiology, ed. W. R.<br />

Bushnell & A. P. Roelfs. Orlando: Academic Press,<br />

pp. 39 77.<br />

Waid, J. S. (1954). Occurrence of aquatic hyphomycetes<br />

upon the root surface of beech grown in woodland<br />

soil. Transactions of the British Mycological Society, 37,<br />

420 421.<br />

Wakefield, A. E., Peters, S. E., Banerji, S., et al. (1993).<br />

Pneumocystis carinii shows DNA homology with the<br />

us<strong>to</strong>myce<strong>to</strong>us red yeast fungi. Molecular Microbiology,<br />

6, 1903 1911.<br />

Walker, G. M. (1998). Yeast Physiology and Biotechnology.<br />

New York: John Wiley.<br />

Walker, J. (2004). Claviceps phalaridis in Australia:<br />

biology, pathology and taxonomy with a description<br />

of the new genus Cepsiclava (Hypocreales,<br />

Clavicipitaceae). Australasian Plant Pathology, 33,<br />

211 239.<br />

Walker, L. B. (1920). Development of Cyathus fascicularis,<br />

C. striatus, and Crucibulum vulgare. Botanical Gazette,<br />

70, 1 24.


808 REFERENCES<br />

Walker, R. F., West, D. C., McLaughlin, S. B. &<br />

Amundsen, C. C. (1989). Growth, xylem pressure<br />

potential, and nutrient absorption of loblolly pine<br />

on a reclaimed surface mine as affected by an<br />

induced Pisolithus tinc<strong>to</strong>rius infection. Forest Science,<br />

35, 569 581.<br />

Walkey, D. G. A. & Harvey, R. (1966a). Studies of the<br />

ballistics of ascospores. New Phy<strong>to</strong>logist, 65, 59 74.<br />

Walkey, D. G. A. & Harvey, R. (1966b). Spore discharge<br />

rhythms in pyrenomycetes. I. A survey of the<br />

periodicity of spore discharge in pyrenomycetes.<br />

Transactions of the British Mycological Society, 49,<br />

583 592.<br />

Wallace, D. R., MacLeod, C. R., Sullivan, C. R., Tyrell, D.<br />

& De Lyzer, A. J. (1976). Induction of resting spore<br />

germination of En<strong>to</strong>mophthora aphidis by long-day<br />

light conditions. Canadian Journal of Botany, 54,<br />

1410 1418.<br />

Wallander, H. & Söderström, B. (1999). Paxillus. In<br />

Ec<strong>to</strong>mycorrhizal <strong>Fungi</strong>. Key Genera in Profile, ed. J. W. G.<br />

Cairney & S. M. Chambers. Berlin: Springer-Verlag,<br />

pp. 231 52.<br />

Walther, V., Rexer, K. & Kost, G. (2001). The on<strong>to</strong>geny of<br />

the fruit bodies of Mycena stylobates. Mycological<br />

Research, 105, 723 733.<br />

Wanderlei-Silva, D., Ne<strong>to</strong>, E. R. & Hanlin, R. (2003).<br />

Molecular systematics of the Phyllachorales<br />

(Ascomycota, <strong>Fungi</strong>) based on 18S ribosomal DNA<br />

sequences. Brazilian Archives of Biology and Technology,<br />

46, 315 322.<br />

Wang, C., Xing, J., Chin, C.-K. & Peters, J. (2002). Fatty<br />

acids with certain structural characteristics are<br />

potent inhibi<strong>to</strong>rs of germination and inducers of cell<br />

death of powdery mildew spores. Physiological and<br />

Molecular Plant Pathology, 61, 151 161.<br />

Wang, Y. & Hall, I. R. (2004). Edible ec<strong>to</strong>mycorrhizal<br />

mushrooms: challenges and achievements. Canadian<br />

Journal of Botany, 82, 1063 1073.<br />

Wang, Y., Aisen, P. & Casadevall, A. (1995). Cryp<strong>to</strong>coccus<br />

neoformans melanin and virulence: mechanism of<br />

action. Infection and Immunity, 63, 3131 3136.<br />

Wang, Z., Binder, M., Dai, Y.-C. & Hibbett, D. S. (2004).<br />

Phylogenetic relationships of Sparassis inferred from<br />

nuclear and mi<strong>to</strong>chondrial ribosomal DNA and RNA<br />

polymerase sequences. Mycologia, 96, 1015 1029.<br />

Warcup, J. H. & Talbot, P. H. B. (1962). Ecology and<br />

identity of mycelia isolated from soil. Transactions of<br />

the British Mycological Society, 45, 495 518.<br />

Warcup, J. H. & Talbot, P. H. B. (1966). Perfect states of<br />

some rhizoc<strong>to</strong>nias. Transactions of the British<br />

Mycological Society, 49, 427 435.<br />

Ward, E. & Adams, M. J. (1998). Analysis of ribosomal<br />

DNA sequences of Polymyxa species and related fungi<br />

and the development of genus- and species-specific<br />

PCR primers. Mycological Research, 102, 965 974.<br />

Ward, H. M. (1882). Researches on the life-his<strong>to</strong>ry of<br />

Hemileia vastatrix. Journal of the Linnean Society, 19,<br />

299 335.<br />

Ward, H. M. (1888). A lily disease. Annals of Botany, 2,<br />

319 382.<br />

Warren, A. C. & Colhoun, J. (1975). Viability of<br />

sporangia of Phy<strong>to</strong>phthora infestans in relation <strong>to</strong><br />

drying. Transactions of the British Mycological Society, 64,<br />

73 78.<br />

Wasser, S. P. (2002). Medicinal mushrooms as a source<br />

of antitumor and immunomodulating polysaccharides.<br />

Applied Microbiology and Biotechnology, 60,<br />

258 274.<br />

Wasson, R. G. (1968). Soma: Divine Mushroom of<br />

Immortality. New York: Harcourt Brace Jovanovich.<br />

Wastie, R. L. (1991). Breeding for resistance. In<br />

Phy<strong>to</strong>phthora infestans, the Cause of Late Blight of Pota<strong>to</strong>,<br />

ed. D. S. Ingram & P. H. Williams. London: Academic<br />

Press, pp. 193 224.<br />

Watanabe, T. (1994). Two new species of homothallic<br />

Mucor in Japan. Mycologia, 86, 691 695.<br />

Watanabe, T. (1997). Stimulation of perithecium and<br />

ascospore production in Sordaria fimicola by<br />

Armillaria and various fungal species. Mycological<br />

Research, 101, 1190 1194.<br />

Watanabe, T., Watanabe, Y., Fukatsu, T. & Kurane, R.<br />

(2001). Mortierella tsukubaensis sp. nov. from Japan,<br />

with a key <strong>to</strong> the homothallic species. Mycological<br />

Research, 105, 506 509.<br />

Waterhouse, G. M. (1963). Key <strong>to</strong> the species of<br />

Phy<strong>to</strong>phthora de Bary. Mycological Papers, 92, 1 22.<br />

Kew: Commonwealth Mycological Institute.<br />

Waterhouse, G. M. (1967). Key <strong>to</strong> Pythium Pringsheim.<br />

Mycological Papers, 109, 1 15. Kew: Commonwealth<br />

Mycological Institute.<br />

Waterhouse, G. M. (1968). The genus Pythium<br />

Pringsheim. Diagnoses (or descriptions) and figures<br />

from the original papers. Mycological Papers, 110,<br />

1 71. Kew: Commonwealth Mycological Institute.<br />

Waterhouse, G. M. (1970). The genus Phy<strong>to</strong>phthora de<br />

Bary. Diagnoses (or descriptions) and figures from<br />

the original papers. Mycological Papers, 120, 1 59.<br />

Kew: Commonwealth Mycological Institute.<br />

Waterhouse, G. M. (1973). Plasmodiophoromycetes. In<br />

The <strong>Fungi</strong>, An Advanced Treatise IVB: A Taxonomic Review<br />

with Keys, ed. G. C. Ainsworth, F. K. Sparrow &<br />

A. S. Sussman. New York: Academic Press, pp. 75 82.


REFERENCES<br />

809<br />

Waters, H., Butler, R. D. & Moore, D. (1975a). Structure<br />

of aerial and submerged sclerotia of Coprinus lagopus.<br />

New Phy<strong>to</strong>logist, 74, 199 205.<br />

Waters, H., Butler, R. D. & Moore, D. (1975b).<br />

Morphogenesis of aerial sclerotia of Coprinus lagopus.<br />

New Phy<strong>to</strong>logist, 74, 207 213.<br />

Waters, S. D. & Callaghan, A. A. (1999). Conidium<br />

germination in co-occurring Conidiobolus and<br />

Basidiobolus in relation <strong>to</strong> their ecology. Mycological<br />

Research, 103, 1259 1269.<br />

Watkinson, S. C. (1979). Growth of rhizomorphs,<br />

mycelial strands, coremia and sclerotia. In Fungal<br />

Walls and Hyphal Growth, ed. J. H. Burnett & A. P. J.<br />

Trinci. Cambridge: Cambridge University Press,<br />

pp. 93 113.<br />

Watling, R. (1971). Polymorphism in Psilocybe merdaria.<br />

New Phy<strong>to</strong>logist, 70, 307 326.<br />

Watling, R. & Harper, D. B. (1998). Chloromethane<br />

production by wood-rotting fungi and an estimate of<br />

the global flux <strong>to</strong> the atmosphere. Mycological<br />

Research, 102, 769 787.<br />

Watson, A. K., Gressel, J., Sharon, A. & Dinoor, A.<br />

(2000). Colle<strong>to</strong>trichum strains for weed control. In<br />

Colle<strong>to</strong>trichum. Host Specificity, Pathology, and Host-<br />

Pathogen Interaction, ed. D. Prusky, S. Freeman &<br />

M. B. Dickman. St. Paul, USA: APS Press,<br />

pp. 245 265.<br />

Watson, I. A. & Luig, N. H. (1958). Somatic hybridization<br />

in Puccinia graminis var. tritici. Proceedings of the<br />

Linnean Society of New South Wales, 83, 190 195.<br />

Webb, J. & Theodorou, M. K. (1988). A rumen anaerobic<br />

fungus of the genus Neocallimastix: ultrastructure of<br />

the polyflagellate zoospore and young thallus.<br />

BioSystems, 21, 393 401.<br />

Webb, J. & Theodorou, M. K. (1991). Neocallimastix<br />

hurleyensis an anaerobic fungus from the<br />

bovine rumen. Canadian Journal of Botany, 69,<br />

1220 1224.<br />

Webber, J. F. & Brasier, C. M. (1984). The transmission of<br />

Dutch elm disease: a study of the processes involved.<br />

In Invertebrate Microbial Interactions, ed. J. M.<br />

Anderson, A. D. M. Rayner & D. W. H. Wal<strong>to</strong>n.<br />

Cambridge: Cambridge University Press,<br />

pp. 271 306.<br />

Webber, J. F. & Gibbs, J. N. (1989). Insect dissemination<br />

of fungal pathogens of trees. In Insect Fungus<br />

Interactions, ed. N. Wilding, N. M. Collins, P. M.<br />

Hammond & J. F. Webber. London: Academic Press,<br />

pp. 161 93.<br />

Weber, R. W. S. (2002). Vacuoles and the fungal lifestyle.<br />

Mycologist, 16, 10 20.<br />

Weber, R. W. S. & Davoli, P. (2003). Teaching techniques<br />

for mycology: 20. Astaxanthin, a carotenoid of<br />

biotechnological importance from yeast and salmonid<br />

fish. Mycologist, 17, 30 34.<br />

Weber, R. W. S. & Pitt, D. (2001). Filamen<strong>to</strong>us fungi<br />

growth and physiology. In Applied Mycology and<br />

Biotechnology 1: Agriculture and Food Production, ed.<br />

G. G. Khacha<strong>to</strong>urians & D. K. Arora. Amsterdam:<br />

Elsevier, pp. 13 54.<br />

Weber, R. W. S. & Tribe, H. T. (2003). Oil as a substrate<br />

for Mortierella species. Mycologist, 17, 134 139.<br />

Weber, R. W. S. & Webster, J. (1998a). Teaching techniques<br />

for mycology: 5. Basidiobolus ranarum. Mycologist,<br />

12, 148 150.<br />

Weber, R. W. S. & Webster, J. (1998b). Stimulation of<br />

growth and reproduction of Sphaeronaemella fimicola<br />

by other coprophilous fungi. Mycological Research,<br />

102, 1055 1061.<br />

Weber, R. W. S. & Webster, J. (2000a). Teaching techniques<br />

for mycology: 9. Olpidium and Rhizophlyctis.<br />

Mycologist, 14, 17 20.<br />

Weber, R. W. S. & Webster, J. (2000b). Teaching techniques<br />

for mycology: 12. A demonstration of the<br />

teleomorph anamorph connection using Pleospora<br />

herbarum. Mycologist, 14, 171 173.<br />

Weber, R. W. S. & Webster, J. (2001a). Teaching techniques<br />

for mycology: 13. Functioning of cleis<strong>to</strong>thecia<br />

in Phyllactinia guttata. Mycologist, 15, 26 30.<br />

Weber, R. W. S. & Webster, J. (2001b). Teaching techniques<br />

for mycology: 14. Mycorrhizal infection of<br />

orchid seedlings in the labora<strong>to</strong>ry. Mycologist, 15,<br />

55 59.<br />

Weber, R. W. S. & Webster, J. (2003). Teaching techniques<br />

for mycology: 21. Sclerotinia, Botrytis and Monilia<br />

(Ascomycota, Leotiales, Sclerotiniaceae). Mycologist,<br />

17, 111 115.<br />

Weber, R. W. S., Wakley, G. E. & Pitt, D. (1998).<br />

His<strong>to</strong>chemical and ultrastructural characterization<br />

of fungal mi<strong>to</strong>chondria. Mycologist, 12, 174 179.<br />

Weber, R. W. S., Webster, J., Barnes, J. C. & Pitt, D. (1999).<br />

Teaching techniques for mycology: 7. Zoospore<br />

discharge by Pythium and Phy<strong>to</strong>phthora. Mycologist, 13,<br />

117 119.<br />

Weber, R. W. S., Wakley, G. E., Thines, E. & Talbot, N. J.<br />

(2001). The vacuole as central element of the lytic<br />

system and sink for lipid droplets in maturing<br />

appressoria of Magnaporthe grisea. Pro<strong>to</strong>plasma, 216,<br />

101 112.<br />

Webster, J. (1952). Spore projection in the<br />

Hyphomycete Nigrospora sphaerica. New Phy<strong>to</strong>logist, 51,<br />

229 235.


810 REFERENCES<br />

Webster, J. (1964). Culture studies on Hypocrea and<br />

Trichoderma species. I. Comparison of the perfect and<br />

imperfect states of Hypocrea gelatinosa, H. rufa and<br />

Hypocrea sp. 1. Transactions of the British Mycological<br />

Society, 47, 75 96.<br />

Webster, J. (1966). Spore projection in Epicoccum and<br />

Arthrinium. Transactions of the British Mycological Society,<br />

49, 339 343.<br />

Webster, J. (1980). <strong>Introduction</strong> <strong>to</strong> <strong>Fungi</strong>, 2nd edn.<br />

Cambridge: Cambridge University Press.<br />

Webster, J. (1987). Convergent evolution and the<br />

functional significance of spore shape in aquatic and<br />

semi-aquatic fungi. In Evolutionary Biology of the <strong>Fungi</strong>,<br />

ed. A. D. M. Rayner, C. M. Brasier & D. Moore.<br />

Cambridge: Cambridge University Press,<br />

pp. 191 201.<br />

Webster, J. (1992). Anamorph teleomorph relationships.<br />

In The Ecology of Aquatic Hyphomycetes, ed.<br />

F. Bärlocher. Berlin: Springer-Verlag, pp. 99 117.<br />

Webster, J., ed. (2006a). Mycology Vol. 1. Interactive DVD-<br />

ROM. Göttingen: IWF Wissen und Medien.<br />

Webster, J., ed. (2006b). Mycology Vol. 2. Interactive DVD-<br />

ROM. Göttingen: IWF Wissen und Medien.<br />

Webster, J. & Chien, C.-Y. (1990). Ballis<strong>to</strong>spore<br />

discharge. Transactions of the Mycological Society of<br />

Japan, 31, 301 305.<br />

Webster, J. & Davey, R. A. (1984). Sigmoid conidial<br />

shape in aquatic fungi. Transactions of the British<br />

Mycological Society, 83, 43 52.<br />

Webster, J. & Dennis, C. (1967). The mechanism of<br />

sporangial discharge in Pythium middle<strong>to</strong>nii. New<br />

Phy<strong>to</strong>logist, 66, 307 313.<br />

Webster, J. & Descals, E. (1981). Morphology, distribution<br />

and ecology of conidial fungi in fresh-water<br />

habitats. In Biology of Conidial <strong>Fungi</strong> 1, ed. G. T. Cole &<br />

B. Kendrick. New York: Academic Press, pp. 295 355.<br />

Webster, J. & Hard, T. (1998a). Alternation of generations<br />

in Allomyces (Blas<strong>to</strong>cladiales). VHS Videotape<br />

2656. Göttingen: Institut für den Wissenschaftlichen<br />

Film.<br />

Webster, J. & Hard, T. (1998b). Ballis<strong>to</strong>spore Discharge in<br />

Basidiomycetes. VHS Videotape C1993. Göttingen:<br />

Institut für den Wissenschaftlichen Film.<br />

Webster, J. & Hard, T. (1999). Pilobolus, a specialized<br />

coprophilous fungus. VHS Videotape C1954.<br />

Göttingen: Institut für den Wissenschaftlichen Film.<br />

Webster, J. & Towfik, F. H. (1972). Sporulation of aquatic<br />

hyphomycetes in relation <strong>to</strong> aeration. Transactions<br />

of the British Mycological Society, 59, 353 364.<br />

Webster, J. & Weber, R. W. S. (1997). Teaching techniques<br />

for mycology: 1. The bird’s nest fungus, Cyathus<br />

stercoreus. Mycologist, 11, 103 105.<br />

Webster, J. & Weber, R. W. S. (1999). Teaching techniques<br />

for mycology: 8. Fruiting cultures of<br />

Sphaerobolus stellatus. Mycologist, 13, 151 153.<br />

Webster, J. & Weber, R. W. S. (2000). Rhizomorphs and<br />

perithecial stromata of Podosordaria tulasnei.<br />

Mycologist, 14, 41 44.<br />

Webster, J. & Weber, R. W. S. (2001). Teaching techniques<br />

for mycology: 15. Fertilization and apothecium<br />

development in Pyronema domesticum and Ascobolus<br />

furfuraceus. Mycologist, 15, 126 131.<br />

Webster, J. & Weber, R. W. S. (2004). Teaching techniques<br />

for mycology: 23. Eclosion of Hypoxylon fragiforme<br />

ascospores as a prelude <strong>to</strong> germination.<br />

Mycologist, 18, 170 173.<br />

Webster, J., Rifai, M. A. & El-Abyad, M. S. (1964). Culture<br />

observations on some discomycetes from burnt<br />

ground. Transactions of the British Mycological Society,<br />

47, 445 454.<br />

Webster, J., Sanders, P. F. & Descals, E. (1978).<br />

Tetraradiate propagules in two species of<br />

En<strong>to</strong>mophthora. Transactions of the British Mycological<br />

Society, 70, 472 479.<br />

Webster, J., Davey, R. A., Duller, G. A. & Ingold, C. T.<br />

(1984a). Ballis<strong>to</strong>spore discharge in Itersonilia perplexans.<br />

Transactions of the British Mycological Society, 82,<br />

13 29.<br />

Webster, J., Davey, R. A. & Ingold, C. T. (1984b). Origin of<br />

the liquid in Buller’s drop. Transactions of the British<br />

Mycological Society, 83, 524 527.<br />

Webster, J., Proc<strong>to</strong>r, M. C. F., Davey, R. A. & Duller, G. A.<br />

(1988). Measurement of the electrical charge on<br />

some basidiospores and an assessment of two<br />

possible mechanisms of basidiospore propulsion.<br />

Transactions of the British Mycological Society, 91,<br />

193 203.<br />

Webster, J., Davey, R. A. & Turner, J. C. R. (1989). Vapour<br />

as the source of water in Buller’s drop. Mycological<br />

Research, 93, 297 302.<br />

Webster, J., Davey, R. A., Smirnoff, N., et al. (1995).<br />

Manni<strong>to</strong>l and hexoses are components of Buller’s<br />

drop. Mycological Research, 99, 833 838.<br />

Webster, J., Pitt, D., Barnes, J. C. & Weber, R. W. S. (1999).<br />

Teaching techniques for mycology: 6. Puccinia graminis,<br />

cause of black stem rust. Mycologist, 13, 15 18.<br />

Webster, M. A. & Dixon, G. R. (1991). Boron, pH and<br />

inoculum concentration influencing colonization by<br />

Plasmodiophora brassicae. Mycological Research, 95,<br />

74 79.<br />

Wedin, M., Wiklund, E., Crewe, A., et al. (2005).<br />

Phylogenetic relationships of Lecanoromycetes (Ascomycota)<br />

as revealed by analyses of mtSSU and nLSU rDNA<br />

sequence data. Mycological Research, 109, 159 172.


REFERENCES<br />

811<br />

Weeks, R. J., Padhye, A. A. & Ajello, L. (1985). His<strong>to</strong>plasma<br />

capsulatum variety farciminosum: a new<br />

combination for His<strong>to</strong>plasma farciminosum. Mycologia,<br />

77, 964 970.<br />

Wehmeyer, L. E. (1926). A biologic and phylogenetic<br />

study of stromatic Sphaeriales. American Journal of<br />

Botany, 13, 575 645.<br />

Wehmeyer, L. E. (1955). Development of Pleospora<br />

armeriae of the Pleospora herbarum complex. Mycologia,<br />

47, 821 834.<br />

Weiergang, I., Lyngs Jørgensen, H. J., Møller, I. M., Friis,<br />

P. & Smedegaard-Petersen, V. (2002). Correlation<br />

between sensitivity of barley <strong>to</strong> Pyrenophora teres<br />

<strong>to</strong>xins and susceptibility <strong>to</strong> the fungus. Physiological<br />

and Molecular Plant Pathology, 60, 121 129.<br />

Weijer, C. J. (2004). Dictyostelium morphogenesis. Current<br />

Opinion in Genetics and Development, 14, 392 398.<br />

Weiler, K. S. & Broach, J. R. (1992). Donor locus selection<br />

during Saccharomyces cerevisiae mating type interconversion<br />

responds <strong>to</strong> distal regula<strong>to</strong>ry signals. Genetics,<br />

132, 929 942.<br />

Weir, A. & Blackwell, M. (2001). Molecular data support<br />

the Laboulbeniales as a separate class of Ascomycota,<br />

the Laboulbeniomycetes. Mycological Research, 105,<br />

1182 1190.<br />

Weiss, M. & Oberwinkler, F. (2001). Phylogenetic<br />

relationships in Auriculariales and related groups<br />

hypotheses derived from nuclear ribosomal DNA<br />

sequences. Mycological Research, 105, 403 415.<br />

Weitzman, I. & Summerbell, R. C. (1995). The derma<strong>to</strong>phytes.<br />

Clinical Microbiology Reviews, 8, 240 259.<br />

Welch, A. M. & Bultmann, T. L. (1993). Natural release of<br />

Epichloe typhina ascospores and its temporal relationship<br />

<strong>to</strong> fly parasitism. Mycologia, 85, 756 763.<br />

Weller, D. M., Raajmakers, J. M., McSpadden Gardener,<br />

B. B. & Thomashow, L. S. (2002). Microbial populations<br />

responsible for specific soil suppressiveness <strong>to</strong><br />

plant pathogens. Annual Review of Phy<strong>to</strong>pathology, 40,<br />

309 348.<br />

Wells, K. (1965). Ultrastructural features of developing<br />

and mature basidia of Schizophyllum commune.<br />

Mycologia, 57, 236 261.<br />

Wells, K. (1970). Light and electron microscopic studies<br />

of Ascobolus stercorarius. I. Nuclear divisions in the<br />

ascus. Mycologia, 62, 761 790.<br />

Wells, K. (1972). Light and electron microscopic studies<br />

of Ascobolus stercorarius. II. Ascus and ascospore<br />

on<strong>to</strong>geny. University of California Publications in Botany,<br />

62, 1 93.<br />

Wells, K. (1987). Comparative morphology, intracompatibility,<br />

and interincompatibility of several species<br />

of Exidiopsis (Exidiaceae). Mycologia, 79, 274 288.<br />

Wells, K. (1994). Jelly fungi, then and now! Mycologia,<br />

86, 18 48.<br />

Wells, K. & Bandoni, R. J. (2001). Heterobasidiomycetes.<br />

In The Mycota VIIB: Systematics and Evolution, ed.<br />

D. J. McLaughlin, E. G. McLaughlin & P. A. Lemke.<br />

Berlin: Springer-Verlag, pp. 85 120.<br />

Weresub, L. K. & LeClair, P. M. (1971). On Papulaspora<br />

and bulbilliferous basidiomycetes Burgoa and<br />

Minimedusa. Canadian Journal of Botany, 49,<br />

2203 2213.<br />

Wessels, J. G. H. (1993a). Wall growth, protein excretion<br />

and morphogenesis in fungi. New Phy<strong>to</strong>logist, 123,<br />

397 413.<br />

Wessels, J. G. H. (1993b). Fruiting in the higher fungi.<br />

Advances in Microbial Physiology, 34, 147 202.<br />

Wessels, J. G. H. (1994). Development of fruit bodies in<br />

Homobasidiomycetes. In The Mycota I: Growth,<br />

Differentiation and Sexuality, ed. J. G. H. Wessels &<br />

F. Meinhardt. Berlin: Springer-Verlag, pp. 351 66.<br />

Wessels, J. G. H. (1997). Hydrophobins: proteins that<br />

change the nature of the fungal surface. Advances in<br />

Microbial Physiology, 38, 1 45.<br />

Wessels, J. G. H. (2000). Hydrophobins, unique fungal<br />

proteins. Mycologist, 14, 153 159.<br />

Wessels, J. G. H. & Sietsma, J. H. (1981). Fungal cell walls:<br />

a survey. In Encyclopedia of Plant Physiology, New Series<br />

13B, ed. W. Tanner & F. A. Loewus. New York:<br />

Springer Verlag, pp. 352 94.<br />

Wessels, J. G. H., Kreger, D. R., Marchant, R.,<br />

Regensburg, B. A. & de Vries, O. M. H. (1972).<br />

Chemical and morphological characterization<br />

of the hyphal wall surface of the basidiomycete<br />

Schizophyllum commune. Biochimica et Biophysica Acta,<br />

273, 346 358.<br />

Wessels, J. G. H., Mol, P. C., Sietsma, J. H. & Vermeulen,<br />

C. A. (1990). Wall structure, wall growth and fungal<br />

cell morphogenesis. In Biochemistry of Cell Walls and<br />

Membranes in <strong>Fungi</strong>, ed. P. J. Kuhn, A. P. J. Trinci, M. J.<br />

Jung, M. W. Goosey & L. G. Copping. Berlin: Springer-<br />

Verlag, pp. 81 95.<br />

West, J. S. (2000). Chemical control of Armillaria. In<br />

Armillaria Root Rot: Biology and Control of Honey Fungus,<br />

ed. R. T. V. Fox. Andover, UK: Intercept Ltd,<br />

pp. 173 82.<br />

Western, J. H. & Cavett, J. J. (1959). The choke disease of<br />

cocksfoot (Dactylis glomerata) caused by Epichloe<br />

typhina (Fr.) Tul. Transactions of the British Mycological<br />

Society, 42, 298 307.<br />

Westfall, P. J. & Momany, M. (2002). Aspergillus nidulans<br />

septin AspB plays pre- and postmi<strong>to</strong>tic roles in<br />

septum, branch, and conidiophore development.<br />

Molecular Biology of the Cell, 13, 110 118.


812 REFERENCES<br />

Whalley, A. J. S. (1985). The Xylariaceae: some ecological<br />

considerations. Sydowia, 38, 369 382.<br />

Whalley, A. J. S. (1987). Xylaria inhabiting fallen fruits.<br />

Agarica, 8, 68 72.<br />

Whalley, A. J. S. (1996). Presidential address. The xylariaceous<br />

way of life. Mycological Research, 100,<br />

897 922.<br />

Whalley, A. J. S. & Greenhalgh, G. N. (1973). Numerical<br />

taxonomy of Hypoxylon. II. A key <strong>to</strong> the identification<br />

of the British species of Hypoxylon. Transactions of the<br />

British Mycological Society, 61, 455 459.<br />

Whetzel, H. H. (1945). A synopsis of the genera and<br />

species of the Sclerotiniaceae, a family of<br />

stromatic inoperculate Discomycetes. Mycologia,<br />

37, 648 714.<br />

Whetzel, H. H. (1946). The cypericolous and juncicolous<br />

species of Sclerotinia. Farlowia, 2, 385 437.<br />

Whisler, H. C. (1987). On the isolation and culture of<br />

water molds: the Blas<strong>to</strong>cladiales and<br />

Monoblepharidales. In Zoosporic <strong>Fungi</strong> in Teaching and<br />

Research, ed. M. S. Fuller & A. Jaworski. Athens, USA:<br />

Southeastern Publishing Corporation, pp. 121 124.<br />

Whisler, H. C. & Marek, L. E. (1987). Monoblepharis<br />

macrandra. In Zoosporic <strong>Fungi</strong> in Teaching and Research,<br />

ed. M. S. Fuller & A. Jaworski. Athens, USA:<br />

Southeastern Publishing Corporation, pp. 54 55.<br />

Whisler, H. C., Zebold, S. L. & Shemanchuk, J. A. (1975).<br />

Life cycle of Coelomomyces psophorae. Proceedings of the<br />

National Academy of Sciences of the USA, 72, 693 696.<br />

White, J. A., Calvert, O. H. & Brown, M. F. (1973).<br />

Ultrastructure of the conidia of Helminthosporium<br />

maydis. Canadian Journal of Botany, 51, 2006 2008.<br />

White, J. F. (1988). Endophyte host associations in<br />

forage grasses. XI. A proposal concerning origin and<br />

evolution. Mycologia, 80, 442 446.<br />

White, J. F. (1993). Endophyte host associations in<br />

forage grasses. XIX. A systematic study of some<br />

sympatric species of Epichloe in England. Mycologia,<br />

85, 444 455.<br />

White, J. F. (1997). Perithecial on<strong>to</strong>geny in the fungal<br />

genus Epichloe: an examination of the clavicipitalean<br />

centrum. American Journal of Botany, 84, 170 178.<br />

White, J. F. & Bultman, T. L. (1987). Endophyte host<br />

associations in forage grasses. VIII. Heterothallism in<br />

Epichloë typhina. American Journal of Botany, 74,<br />

1716 1721.<br />

White, J. F., Morrow, A. C., Morgan-Jones, G. &<br />

Chambless, D. A. (1991). Endophyte host associations<br />

in forage grasses. XIV. Primary stromata<br />

formation and seed transmission in Epichloe typhina:<br />

developmental and regula<strong>to</strong>ry aspects. Mycologia, 83,<br />

72 81.<br />

White, J. F., Martin, T. I. & Cabral, D. (1996). Endophyte<br />

associations in grasses. XXII. Conidia formation by<br />

Acremonium endophytes on the phylloplanes of<br />

Agrostis hiemalis and Poa rigidifolia. Mycologia, 88,<br />

174 178.<br />

White, T. J., Bruns, T., Lee, S. & Taylor, J. (1990).<br />

Amplification and direct sequencing of fungal<br />

ribosomal RNA genes for phylogenetics. In PCR<br />

Pro<strong>to</strong>cols: A Guide <strong>to</strong> Methods and Applications, ed.<br />

M. A. Innis, D. H. Gelfand, J. J. Sninsky & T. J. White.<br />

New York: Academic Press, pp. 315 22.<br />

Whiteside, W. C. (1957). Perithecial initials of<br />

Chae<strong>to</strong>mium. Mycologia, 49, 420 425.<br />

Whiteside, W. C. (1961). Morphological studies in the<br />

Chae<strong>to</strong>miaceae. I. Mycologia, 53, 512 523.<br />

Whitney, K. D. & Arnott, H. J. (1986). Morphology and<br />

development of calcium oxalate deposits<br />

in Gilbertella persicaria (Mucorales). Mycologia, 78,<br />

42 51.<br />

Whittaker, R. H. (1969). New concepts of kingdoms of<br />

organisms. Science, 163, 150 160.<br />

Wickes, B. L., Mayorga, M. E., Edman, U. & Edman, J. C.<br />

(1996). Dimorphism and haploid fruiting in<br />

Cryp<strong>to</strong>coccus neoformans: association with the a-<br />

mating type. Proceedings of the National Academy of<br />

Sciences of the USA, 93, 7327 7331.<br />

Wicklow, D. T. (1979). Hair ornamentation and preda<strong>to</strong>r<br />

defence in Chae<strong>to</strong>mium. Transactions of the British<br />

Mycological Society, 72, 107 110.<br />

Wicklow, D. T., Langie, R., Crabtree, S. & Detroy, R. W.<br />

(1984). Degradation of lignocellulose in wheat straw<br />

versus hardwood by Cyathus and related species<br />

(Nidulariaceae). Canadian Journal of Microbiology, 30,<br />

632 636.<br />

Wieland, T. (1996). Toxins and psychoactive<br />

compounds from mushrooms. In The Mycota VI:<br />

Human and Animal Relationships, ed. D. H. Howard &<br />

J. D. Miller. Berlin: Springer-Verlag, pp. 229 48.<br />

Wieland, T. & Faulstich, H. (1991). Fifty years of<br />

amanitin. Experientia, 47, 1186 1193.<br />

Wilding, N. & Lauckner, F. B. (1974). En<strong>to</strong>mophthora<br />

infecting wheat bulb fly at Rothamsted,<br />

Hertfordshire. Annals of Applied Biology, 76, 161 170.<br />

Willetts, H. J. (1969). Structure of the outer surfaces of<br />

sclerotia of certain fungi. Archiv für Mikrobiologie, 69,<br />

48 53.<br />

Willetts, H. J. (1971). The survival of fungal sclerotia<br />

under adverse environmental conditions. Biological<br />

Reviews, 46, 387 407.<br />

Willetts, H. J. (1972). The morphogenesis and possible<br />

evolutionary origins of fungal sclerotia. Biological<br />

Reviews, 47, 515 536.


REFERENCES<br />

813<br />

Willetts, H. J. & Bullock, S. (1992).<br />

Developmental biology of sclerotia. Mycological<br />

Research, 96, 801 816.<br />

Willetts, H. J. & Wong, J. A.-L. (1980). The biology of<br />

Sclerotinia sclerotiorum, S. trifoliorum, and S. minor with<br />

emphasis on specific nomenclature. Botanical Review,<br />

46, 101 165.<br />

Williams, E. N. D., Todd, N. K. & Rayner, A. D. M. (1981).<br />

Spatial development of populations of Coriolus<br />

versicolor. New Phy<strong>to</strong>logist, 89, 307 319.<br />

Williams, M. A. J., Beckett, A. & Read, N. D. (1985).<br />

Ultrastructural aspects of fruit body differentiation<br />

in Flammulina velutipes. InDevelopmental Biology of<br />

Higher <strong>Fungi</strong>, ed. D. Moore, L. A. Cassel<strong>to</strong>n, D. A. Wood<br />

& J. C. Frankland. Cambridge: Cambridge University<br />

Press, pp. 429 50.<br />

Williams, P. G. (1984). Obligate parasitism and axenic<br />

culture. In The Cereal Rusts 1: Origins, Specificity,<br />

Structure, and Physiology, ed. W. R. Bushnell &<br />

A. P. Roelfs. Orlando: Academic Press, pp. 399 430.<br />

Williams, P. G., Scott, K. J. & Kuhl, J. L. (1966). Vegetative<br />

growth of Puccinia graminis f. sp. tritici in vitro.<br />

Phy<strong>to</strong>pathology, 56, 1418 1419.<br />

Williams, R. H. & Fitt, B. D. L. (1999). Differentiating<br />

A and B groups of Lep<strong>to</strong>sphaeria maculans, causal<br />

agent of stem canker (blackleg) of oilseed rape. Plant<br />

Pathology, 48, 161 175.<br />

Williams, R. H., Whipps, J. M. & Cooke, R. C. (1998). Role<br />

of soil mesofauna in dispersal of Coniothyrium<br />

minitans: mechanisms of transmission. Soil Biology<br />

and Biochemistry, 30, 1937 1945.<br />

Williams, R. J. (1984). Downy mildews of tropical<br />

cereals. Advances in Plant Pathology, 3, 1 103.<br />

Willocquet, L. & Clerjeau, M. (1998). An analysis of the<br />

effects of environmental fac<strong>to</strong>rs on conidial dispersal<br />

of Uncinula neca<strong>to</strong>r (grape powdery mildew) in<br />

vineyards. Plant Pathology, 47, 227 233.<br />

Willoughby, L. G. (1962). The fruiting behaviour and<br />

nutrition of Cladochytrium replicatum Karling. Annals<br />

of Botany N.S., 26, 13 36.<br />

Willoughby, L. G. (1994). <strong>Fungi</strong> and Fish Diseases. Stirling:<br />

Pisces Press.<br />

Willoughby, L. G. (1998a). Saprolegnia polymorpha sp.<br />

nov., a fungal parasite on Koi carp in the U.K. Nova<br />

Hedwigia, 66, 507 511.<br />

Willoughby, L. G. (1998b). A quantitative ecological<br />

study on the monocentric soil chytrid, Rhizophlyctis<br />

rosea, in Provence. Mycological Research, 102,<br />

1338 1342.<br />

Willoughby, L. G. (2001). The activity of Rhizophlyctis<br />

rosea in soil: some deductions from labora<strong>to</strong>ry<br />

observations. Mycologist, 15, 113 117.<br />

Willoughby, L. G. & Pickering, A. D. (1977). Viable<br />

Saprolegniaceae spores on the epidermis of the<br />

salmonid fish Salmo trutta and Salvelinus<br />

alpinus. Transactions of the British Mycological Society,<br />

68, 91 95.<br />

Willoughby, L. G. & Roberts, R. J. (1992). Towards<br />

strategic use of fungicides against Saprolegnia parasitica<br />

in salmonid fish hatcheries. Journal of Fish<br />

Diseases, 15, 1 13.<br />

Wilson, C. M. (1952). Meiosis in Allomyces. Bulletin of the<br />

Torrey Botanical Club, 79, 139 160.<br />

Wilson, C. M. & Flanagan, P. W. (1968). The life cycle<br />

and cy<strong>to</strong>logy of Brachyallomyces. Canadian Journal of<br />

Botany, 46, 1361 1367.<br />

Wilson, I. M. (1952). The ascogenous hyphae of Pyronema<br />

confluens. Annals of Botany N.S., 16, 321 339.<br />

Wilson, J. P. & Hanna, W. W. (1998). Smut resistance<br />

and grain yield of pearl millet hybrids near isogenic<br />

at the Tr locus. Crop Science, 38, 649 651.<br />

Wilson, M. & Henderson, D. M. (1966). British Rust <strong>Fungi</strong>.<br />

Cambridge: Cambridge University Press.<br />

Wilson, R. W. & Beneke, E. S. (1966). Basidiospore<br />

germination in Calvatia gigantea. Mycologia, 58,<br />

328 332.<br />

Wingfield, B. D., Ericson, L., Szaro, T. & Burdon, J. J.<br />

(2004). Phylogenetic patterns in the Uredinales.<br />

Australasian Plant Pathology, 33, 327 335.<br />

Wingfield, M. J., Kendrick, W. B. & Schalk van Wyk, P.<br />

(1991). Analysis of conidium on<strong>to</strong>geny in anamorphs<br />

of Ophios<strong>to</strong>ma: Pesotum and Phialographium are synonyms<br />

of Graphium. Mycological Research, 95,<br />

1328 1333.<br />

Winka, K., Eriksson, O. E. & Bång, A. (1998). Molecular<br />

evidence for recognizing the Chae<strong>to</strong>thyriales.<br />

Mycologia, 90, 822 830.<br />

Winner, M., Giménez, A., Schmidt, H., et al. (2004).<br />

Unusual pulvinic acid dimers from the common<br />

fungi Scleroderma citrinum (common earthball) and<br />

Chalciporus piperatus (peppery bolete).<br />

Angewandte Chemie International <strong>Edition</strong>,<br />

43, 1883 1886.<br />

Winterstein, D. (2001). Cordyceps (Kernkeulen),<br />

Gourmets in Pilzreich. Der Tintling, 6, 24 30.<br />

Win-Tin, & Dick, M. W. (1975). Cy<strong>to</strong>logy of Oomycetes.<br />

Evidence for meiosis and multiple chromosome<br />

associations in Saprolegniaceae and Pythiaceae, with<br />

an introduction <strong>to</strong> the cy<strong>to</strong>taxonomy of Achlya and<br />

Pythium. Archives of Microbiology, 105, 283 293.<br />

Wirth, V. (1995a). Die Flechten Baden-Württembergs Teil 1.<br />

Stuttgart: Eugen Ulmer.<br />

Wirth, V. (1995b). Die Flechten Baden-Württembergs Teil 2.<br />

Stuttgart: Eugen Ulmer.


814 REFERENCES<br />

Witthuhn, R. C., Wingfield, B. D., Wingfield, M. J.,<br />

Wolfaardt, M. & Harring<strong>to</strong>n, T. C. (1998). Monophyly<br />

of the conifer species in the Cera<strong>to</strong>cystis coerulescens<br />

complex based on DNA sequence data. Mycologia, 90,<br />

96 101.<br />

Woese, C. R. (1987). Bacterial evolution. Microbiological<br />

Reviews, 51, 221 271.<br />

Wolf, F. T. (1981). The biology of En<strong>to</strong>mophthora. Nova<br />

Hedwigia, 35, 553 599.<br />

Wolf, G. (1982). Physiology and biochemistry of spore<br />

germination. In The Rust <strong>Fungi</strong>, ed. K. J. Scott & A. K.<br />

Chakravorty. London: Academic Press, pp. 151 178.<br />

Wolfe, L. M. & Rissler, L. J. (1999). Reproductive consequences<br />

of a gall-inducing fungal pathogen<br />

(Exobasidium vaccinii) onRhododendron calendulaceum<br />

(Ericaceae). Canadian Journal of Botany, 77, 1454 1459.<br />

Wolpert, T. J., Dunkle, L. D. & Cuiffetti, L. M. (2002).<br />

Host-selective <strong>to</strong>xins and avirulence dominants:<br />

what’s in a name? Annual Review of Phy<strong>to</strong>pathology, 40,<br />

251 285.<br />

Wong, G. J. (1993). Mating and fruiting studies of<br />

Auricularia delicata and A. fuscosuccinea. Mycologia, 85,<br />

187 194.<br />

Wong, G. J. & Wells, K. (1987). Comparative morphology,<br />

compatibility, and interfertility of Auricularia<br />

cornea, A. polytricha, and A. tenuis. Mycologia, 79,<br />

847 856.<br />

Wong, G. J., Wells, K. & Bandoni, R. J. (1985).<br />

Interfertility and comparative morphological studies<br />

of Tremella mesenterica. Mycologia, 77, 36 49.<br />

Wong, S., Fares, M. A., Zimmermann, W., Butler, G. &<br />

Wolfe, K. H. (2003). Evidence from comparative<br />

genomics for a complete sexual cycle in the ‘asexual’<br />

pathogenic yeast Candida glabrata. Genome Biology, 4,<br />

R10.<br />

Wood, H. A. & Bozarth, R. F. (1973). Heterokaryon<br />

transfer of viruslike particles and a cy<strong>to</strong>plasmically<br />

inherited determinant in Ustilago maydis.<br />

Phy<strong>to</strong>pathology, 63, 1019 1021.<br />

Wood, S. N. & Cooke, R. C. (1986). Effect of Pip<strong>to</strong>cephalis<br />

species on growth and sporulation of Pilaira anomala.<br />

Transactions of the British Mycological Society, 86,<br />

672 674.<br />

Wood, S. N. & Cooke, R. C. (1987). Nutritional competence<br />

of Pilaira anomala in relation <strong>to</strong> exploitation of<br />

faecal resource units. Transactions of the British<br />

Mycological Society, 88, 247 255.<br />

Woodham-Smith, C. (1962). The Great Hunger. Ireland<br />

1845 1849. London: Hamish Hamil<strong>to</strong>n.<br />

Woods, J. P. (2002). His<strong>to</strong>plasma capsulatum molecular<br />

genetics, pathogenesis, and responsiveness <strong>to</strong> its<br />

environment. Fungal Genetics and Biology, 35, 81 97.<br />

Woodward, R. C. (1927). Studies on Podosphaera leucotricha<br />

(Ell. & Ev.) Salm. I. The mode of perennation.<br />

Transactions of the British Mycological Society, 12,<br />

173 204.<br />

Woronin, M. (1878). Plasmodiophora brassicae. Urheber<br />

der Kohlpflanzen-Hernie. Jahrbücher für<br />

Wissenschaftliche Botanik, 11, 548 574.<br />

Worrall, J. J., Anagnost, S. E. & Zabel, R. A. (1997).<br />

Comparison of wood decay among diverse lignicolous<br />

fungi. Mycologia, 89, 199 219.<br />

Wu, C.-G. & Kimbrough, J. W. (1992). Ultrastructural<br />

studies of ascosporogenesis in Ascobolus immersus.<br />

Mycologia, 84, 459 466.<br />

Wu, C.-G. & Kimbrough, J. W. (1993). Ultrastructure of<br />

ascospore on<strong>to</strong>geny in Aleuria, Oc<strong>to</strong>spora and Pulvinella<br />

(Otideaceae, Pezizales). International Journal of Plant<br />

Sciences, 154, 334 349.<br />

Wu, C.-G. & Kimbrough, J. W. (2001). Ultrastructural<br />

studies of ascosporogenesis in Ascobolus stic<strong>to</strong>ideus<br />

(Pezizales, Ascomycetes). International Journal of Plant<br />

Sciences, 162, 71 102.<br />

Wubah, D. A., Fuller, M. S. & Akin, D. E. (1991). Isolation<br />

of monocentric and polycentric fungi from the<br />

rumen and faeces of cows. Canadian Journal of Botany,<br />

69, 1232 1236.<br />

Wyand, R. A. & Brown, J. K. M. (2003). Genetic and forma<br />

specialis diversity in Blumeria graminis of cereals and<br />

its implications for host pathogen co-evolution.<br />

Molecular Plant Pathology, 4, 187 198.<br />

Wynn, A. R. & Ep<strong>to</strong>n, H. A. S. (1979). Parasitism of<br />

oospores of Phy<strong>to</strong>phthora erythroseptica in soil.<br />

Transactions of the British Mycological Society, 73,<br />

255 259.<br />

Xu, J., Kerrigan, R. W., Callac, P., Horgen, P. A. &<br />

Anderson, J. B. (1997). Genetic structure of natural<br />

populations of Agaricus bisporus, the commercial<br />

but<strong>to</strong>n mushroom. Journal of Heredity, 88, 482 488.<br />

Xu, J., Boyd, C. M., Livings<strong>to</strong>n, E., et al. (1999). Species<br />

and genotypic diversities and similarities of pathogenic<br />

yeasts colonizing women. Journal of Clinical<br />

Microbiology, 37, 3835 3843.<br />

Xu, J., Vilgalys, R. & Mitchell, T. G. (2000). Multiple gene<br />

genealogies reveal recent dispersion and hybridization<br />

in the human pathogenic fungus Cryp<strong>to</strong>coccus<br />

neoformans. Molecular Ecology, 9, 1471 1481.<br />

Xu, J.-R. (2000). MAP kinases in fungal pathogens.<br />

Fungal Genetics and Biology, 31, 137 152.<br />

Xu, J.-R. & Hamer, J. E. (1996). MAP kinase and cAMP<br />

signaling regulate infection structure formation and<br />

pathogenic growth in the rice blast fungus<br />

Magnaporthe grisea. Genes and Development, 10,<br />

2696 2706.


REFERENCES<br />

815<br />

Xu, J.-T. & Mu, C. (1990). The relation between growth<br />

of Gastrodia elata pro<strong>to</strong>corms and fungi. Acta Botanica<br />

Sinica, 32, 26 31.<br />

Xu, X. M. & Robinson, D. J. (2000). Epidemiology of<br />

brown rot (Monilinia fructigena) on apple: infection of<br />

fruits by conidia. Plant Pathology, 49, 201 206.<br />

Yamakazi, Y. & Ootaki, T. (1996). Roles of extracellular<br />

fibrils connecting progametangia in mating of<br />

Phycomyces blakesleeanus. Mycological Research, 100,<br />

984 988.<br />

Yan, A. S. (1998). Contamination sources of Pyronema<br />

domesticum on Chinese cot<strong>to</strong>n-based medical<br />

products. Radiation Physics and Chemistry, 52, 7 9.<br />

Yarrow, D. (1998). Methods for the isolation, maintenance<br />

and identification of yeasts. In The Yeasts: A<br />

Taxonomic Study, 4th edn, ed. C. P. Kurtzman & J. W.<br />

Fell. Amsterdam: Elsevier, pp. 77 100.<br />

Yarwood, C. E. (1941). Diurnal cycle of ascus maturation<br />

of Taphrina deformans. American Journal of Botany, 28,<br />

355 357.<br />

Ye, X. S., Lee, S.-L., Wolkow, T. D., et al. (1999).<br />

Interaction between developmental and cell cycle<br />

regula<strong>to</strong>rs is required for morphogenesis in<br />

Aspergillus nidulans. EMBO Journal, 18, 6994 7001.<br />

Yendol, W. G. & Pashke, J. D. (1965). Pathology of an<br />

En<strong>to</strong>mophthora infection in the Eastern subterranean<br />

termite Reticulitermes flavipes (Kollar). Journal of<br />

Invertebrate Pathology, 7, 414 422.<br />

Yokochi, T., Honda, D., Higashihara, T. & Nakahara, T.<br />

(1998). Optimization of docosahexaenoic acid<br />

production by Schizochytrium limacinum SR21. Applied<br />

Microbiology and Biotechnology, 49, 72 76.<br />

Yoon, K. S. & McLaughlin, D. J. (1979). Formation of the<br />

hilar appendix in basidiospores of Boletus rubinellus.<br />

American Journal of Botany, 66, 870 873.<br />

Yoon, K. S. & McLaughlin, D. J. (1984).<br />

Basidiosporogenesis in Boletus rubinellus. I. Sterigmal<br />

initiation and early spore development. American<br />

Journal of Botany, 71, 80 90.<br />

Yoon, K. S. & McLaughlin, D. J. (1986).<br />

Basidiosporogenesis in Boletus rubinellus. II. Late spore<br />

development. Mycologia, 78, 185 197.<br />

Yoon, S. I., Kim, S. Y., Lim, Y. W. & Jung, H. S. (2003).<br />

Phylogenetic evaluation of stereoid fungi. Journal of<br />

Microbiology and Biotechnology, 13, 406 414.<br />

Yoshikawa, M. & Masago, H. (1977). Effect of cyclic AMP<br />

in catabolite-repressed zoosporangial formation in<br />

Phy<strong>to</strong>phthora capsici. Canadian Journal of Botany, 55,<br />

840 843.<br />

Youatt, J. (1977). Chemical nature of Allomyces walls.<br />

Transactions of the British Mycological Society, 69,<br />

187 190.<br />

Young, E. L. (1943). Studies on Labyrinthula. The etiological<br />

agent of the wasting disease of eelgrass.<br />

American Journal of Botany, 30, 586 593.<br />

Yu, K. W., Suh, H. J., Bae, S. H., et al. (2001). Chemical<br />

properties and physiological activities of stromata of<br />

Cordyceps militaris. Journal of Microbiology and<br />

Biotechnology, 11, 266 274.<br />

Yu, L., Lee, K. K., Ens, K., et al. (1994). Partial characterization<br />

of a Candida albicans fimbrial adhesin.<br />

Infection and Immunity, 62, 2834 2842.<br />

Yu, M. Q. & Ko, W. H. (1996). Mating-type segregation in<br />

Choanephora cucurbitarum following sexual reproduction.<br />

Canadian Journal of Botany, 74, 919 923.<br />

Yu, M. Q. & Ko, W. H. (1999). Azygospore formation<br />

following pro<strong>to</strong>plast fusion in Choanephora cucurbitarum.<br />

Mycological Research, 103, 684 688.<br />

Yuan, G. F. & Yong, S. C. (1984). A new obligate<br />

azygosporic species of Rhizopus. Mycotaxon, 20,<br />

397 400.<br />

Yuan, X., Xiao, S. & Taylor, T. N. (2005). Lichen-like<br />

symbiosis 600 million years ago. Science, 308,<br />

1017 1020.<br />

Yuasa, K. & Hatai, K. (1996). Some biochemical<br />

characteristics of the genera Saprolegnia, Achlya and<br />

Aphanomyces isolated from fishes with fungal infection.<br />

Mycoscience, 37, 477 479.<br />

Yukawa, Y. & Tanaka, S. (1979). Scanning electron<br />

microscope observations on resting sporangia of<br />

Plasmodiophora brassicae in clubroot tissues after<br />

alcohol cracking. Canadian Journal of Botany, 57,<br />

2528 2532.<br />

Yurlova, N. A., de Hoog, G. S. & Gerrits van den Ende,<br />

A. H. (1999). Taxonomy of Aureobasidium and allied<br />

genera. Studies in Mycology, 43, 63 69.<br />

Zadoks, J. C. & Bouwman, J. J. (1985). Epidemiology in<br />

Europe. In The Cereal Rusts 2: Diseases, Distribution,<br />

Epidemiology, and Control, ed. A. P. Roelfs & W. R.<br />

Bushnell. Orlando: Academic Press, pp. 329 369.<br />

Zahari, P. & Ship<strong>to</strong>n, W. A. (1988). Growth and<br />

sporulation responses of Basidiobolus ranarum <strong>to</strong><br />

changes in environmental parameters. Transactions of<br />

the British Mycological Society, 91, 141 148.<br />

Zak, J. (1976). Pure culture synthesis of Pacific Madrone<br />

ectendomycorrhizae. Mycologia, 68, 362 369.<br />

Zambino, P. J. & Szabo, L. J. (1993). Phylogenetic<br />

relationships of selected cereal and grass rusts<br />

based on rDNA sequence analysis. Mycologia, 85,<br />

401 414.<br />

Zambonelli, A., Rivetti, C., Percudani, R. & Ot<strong>to</strong>nello, S.<br />

(2000). TuberKey: a DELTA-based <strong>to</strong>ol for the<br />

description and interactive identification of truffles.<br />

Mycotaxon, 74, 57 76.


816 REFERENCES<br />

Zeigler, R. S. (1998). Recombination in Magnaporthe<br />

grisea. Annual Review of Phy<strong>to</strong>pathology, 36, 249 275.<br />

Zelmer, C. D. & Currah, R. S. (1995). Evidence for a<br />

fungal liaison between Corallorhiza trifida<br />

(Orchidaceae) and Pinus con<strong>to</strong>rta (Pinaceae). Canadian<br />

Journal of Botany, 73, 862 866.<br />

Zelmer, C. D., Cuthbertson, L. & Currah, R. S. (1996).<br />

<strong>Fungi</strong> associated with terrestrial orchid mycorrhizas,<br />

seeds and pro<strong>to</strong>corms. Mycoscience, 37, 439 448.<br />

Zemek, J., Marvanová, L., Kuniak, L. & Kadlecikova, B.<br />

(1985). Hydrolytic enzymes in aquatic hyphomycetes.<br />

Folia Microbiologica, 30, 363 372.<br />

Zentmyer, G. A. (1980). Phy<strong>to</strong>phthora cinnamomi and the<br />

diseases it causes. Monograph No. 10. St. Paul, USA:<br />

APS Press.<br />

Zeyen, R. J., Carver, T. L. W. & Lyngkjaer, M. F. (2002).<br />

Epidermal cell papillae. In The Powdery Mildews. A<br />

Comprehensive Treatise, ed. R. R. Bélanger, W. R.<br />

Bushnell, A. J. Dik & T. L. W. Carver. St. Paul, USA: APS<br />

Press, pp. 107 125.<br />

Zhan, J., Mundt, C. C. & McDonald, B. A. (2001). Using<br />

restriction fragment length polymorphisms <strong>to</strong> assess<br />

temporal variation and estimate the number of<br />

ascospores that initiate epidemics in field populations<br />

of Mycosphaerella graminicola. Phy<strong>to</strong>pathology, 91,<br />

1011 1017.<br />

Zhang, G. & Berbee, M. L. (2001). Pyrenophora phylogenetics<br />

from ITS and glyceraldehyde-3-phosphate<br />

dehydrogenase gene sequences. Mycologia, 93,<br />

1048 1063.<br />

Zhang, L., Baasiri, R. A. & van Alfen, N. K. (1998). Viral<br />

repression of fungal pheromone precursor gene<br />

expression. Molecular and Cellular Biology, 18, 953 959.<br />

Zhang, N. & Blackwell, M. (2001). Molecular phylogeny<br />

of dogwood anthracnose fungus (Discula destructiva)<br />

and the Diaporthales. Mycologia, 93, 355 365.<br />

Zhang, Q., Griffith, J. M. & Grant, B. R. (1992). Role of<br />

phosphatidic acid during differentiation of<br />

Phy<strong>to</strong>phthora palmivora zoospores. Journal of General<br />

Microbiology, 138, 451 459.<br />

Zhang, W., Dick, W. A. & Hoitink, H. A. J. (1996).<br />

Compost-induced systemic acquired resistance in<br />

cucumber <strong>to</strong> Pythium root rot and anthracnose.<br />

Phy<strong>to</strong>pathology, 86, 1066 1070.<br />

Zheng, R. Y. & Chen, G. Q. (2001). A monograph of<br />

Cunninghamella. Mycotaxon, 80, 1 75.<br />

Zhou, F. S., Zhang, Z. G., Gregersen, P. L., et al. (1998).<br />

Molecular characterization of the oxalate oxidase<br />

involved in the response of barley <strong>to</strong> the powdery<br />

mildew fungus. Plant Physiology, 117, 33 41.<br />

Zhou, X.-L., Stumpf, M. A., Hoch, H. C. & Kung, C. (1991).<br />

A mechanosensitive channel in whole cells and in<br />

membrane patches of the fungus Uromyces. Science,<br />

253, 1415 1417.<br />

Zlotnik, H., Fernandez, M. P., Bowers, B. & Cabib, E.<br />

(1984). Saccharomyces cerevisiae mannoproteins form<br />

an external cell wall layer that determines wall<br />

porosity. Journal of Bacteriology, 159, 1018 1026.<br />

Zoberi, M. H. (1985). Propagule liberation in the<br />

Mucorales. Botanical Journal of the Linnean Society, 91,<br />

167 173.<br />

Zugmaier, W., Bauer, R. & Oberwinkler, F. (1994).<br />

Mycoparasitism of some Tremella species. Mycologia,<br />

86, 49 56.<br />

Zuppinger, C. & Roos, U.-P. (1997). Cell shape, motility<br />

and distribution of F-actin in amoebae of the<br />

myce<strong>to</strong>zoans Pro<strong>to</strong>stelium mycophaga and Acrasis rosea.<br />

A comparison with Dictyostelium discoideum. European<br />

Journal of Protis<strong>to</strong>logy, 33, 396 408.<br />

Zureik, M., Neukirch, C., Leynaert, B., et al. (2002).<br />

Sensitisation <strong>to</strong> airborne moulds and severity of<br />

asthma: cross sectional study from European<br />

Community respira<strong>to</strong>ry health survey. British Medical<br />

Journal, 325, 411 414.<br />

Zycha, H. & Siepmann, R. & Linnemann, G. (1969).<br />

Mucorales. Lehre, Germany: J. Cramer.


Index<br />

Page numbers with images are underlined, those with explanations of concepts are printed in bold.<br />

ABC transporters 280, 383, 439<br />

Absidia 183<br />

Absidia corymbifera 184<br />

Absidia glauca 172, 176, 185<br />

Absidia spinosa 173, 176, 184<br />

Acaulospora 221<br />

acervulus 231, 387<br />

Achlya 86–91; aplanetic forms 93;<br />

asexual reproduction 87, 88;<br />

hyphae 80; relative sexuality 91;<br />

sex hormones 88–89, 90, 91; sexual<br />

reproduction 88–91<br />

Achlya ambisexualis 88, 89, 91<br />

Achlya bisexualis 88<br />

Achlya colorata 69, 87–88<br />

Achlya heterosexualis 88<br />

Achlya klebsiana 87<br />

Acrasiomycetes (acrasid cellular<br />

slime moulds) 40–41<br />

Acrasis rosea 41<br />

Acremonium 316, 339, 341<br />

Acremonium chrysogenum 303, 349<br />

Acremonium salmosynnematum 349<br />

actin cap 9<br />

actin filaments 9, 10, 51, 65–66, 158,<br />

258, 271, 646; in ascospore<br />

formation 323; in graviperception<br />

549; in penetration peg 381<br />

actin-myosin 46, 51, 72; see also<br />

cy<strong>to</strong>skele<strong>to</strong>n<br />

actin ring 259, 271, 301<br />

adaxial blob (drop) 493–494, 496<br />

adhesins 277<br />

adhesion pad 615<br />

adhesives 66, 205, 364, 378, 381, 395,<br />

438; in nema<strong>to</strong>phagous fungi 683<br />

adhesorium 59<br />

adnate gills 523<br />

adnexed gills 523<br />

adult plant resistance; see host<br />

resistance<br />

aeciospore 613, Pl. 12<br />

aecium 612, 613, 622, Pl. 12; caeoma<br />

635; roestelioid aecium 632<br />

Aegerita 701<br />

Aegerita candida (teleom. Bulbillomyces<br />

farinosus) 506, 697, 699<br />

Aegeritina 701<br />

Aegeritina <strong>to</strong>rtuosa (teleom.<br />

Subulicystidium longisporum) 697<br />

aequi-hymenial gills 522<br />

aero-aquatic fungi 506, 696–701;<br />

anamorph-teleomorph connections<br />

697; ecophysiology 698–701<br />

aethalium 50, Pl. 1<br />

afla<strong>to</strong>xins 304, 305, Pl. 4<br />

Agaricus 532–536<br />

Agaricus arvensis 532<br />

Agaricus bisporus 15, 533; breeding<br />

536; cultivation 525, 532–534;<br />

life cycle 535; mating system 506,<br />

535; morphogenesis 534–535; var.<br />

burnettii 536; var. eurotetrasporus 536<br />

Agaricus bi<strong>to</strong>rquis 532<br />

Agaricus brunnescens; see Agaricus<br />

bisporus<br />

Agaricus campestris 496, 521, 532;<br />

mating system 506<br />

Agaricus macrosporus 532<br />

Agaricus silvaticus 515, 532<br />

Agaricus tabularis 532<br />

Agaricus xanthodermus 532<br />

age of fungi; see evolution<br />

ageing 271<br />

agglutination 254, 265<br />

aggregation; see Dictyostelium<br />

discoideum<br />

Agrobacterium transformation 536<br />

agroclavine 354<br />

AIDS 279, 306–307, 661–662, 664<br />

air pollution moni<strong>to</strong>ring, lichens 454;<br />

Rhytisma 442; Sporobolomyces roseus<br />

669<br />

Ajellomyces 290<br />

Ajellomyces capsulatus 289;<br />

see His<strong>to</strong>plasma capsulatum<br />

Ajellomyces dermatitidis;<br />

see Blas<strong>to</strong>myces dermatitidis<br />

Ala<strong>to</strong>spora acuminata 685, 687<br />

Albugo 122–125<br />

Albugo bliti 124<br />

Albugo candida 122, 123, 124–125, Pl. 2;<br />

haus<strong>to</strong>rium 122; oospore 124<br />

Albugo tragopogonis 122, 125<br />

alcohol and fungi; see coprine<br />

alcoholic fermentation 254, 262<br />

alder, leaf blister 251<br />

Aleuria 417–419<br />

Aleuria aurantia 243, 419, Pl. 6;<br />

mycorrhiza 419<br />

aleuriaxanthin 419<br />

algal parasites 127<br />

alkaloids 354, 364, 539, 541;<br />

biosynthesis 354; commercial<br />

production 354; see Claviceps<br />

purpurea, Neotyphodium<br />

allergens; see asthma<br />

allyl amines 279, 280<br />

Allomyces 155–160; Brachy-Allomyces 160;<br />

Cys<strong>to</strong>genes 159; Eu-Allomyces 156; life<br />

cycle 158; polyploidy 159;<br />

sex hormones (parisin and sirenin)<br />

156<br />

Allomyces anomalus 155, 160<br />

Allomyces arbuscula 155, 157<br />

Allomyces macrogynus 155, 158<br />

Allomyces moniliformis 155, 159<br />

Allomyces neo-moniliformis 159<br />

Alternaria 469–471; plant diseases 472;<br />

sexual ances<strong>to</strong>rs 469<br />

Alternaria alternata 470<br />

Alternaria brassicae 469, 471<br />

Alternaria brassicicola 469<br />

Alternaria solani 469<br />

Alternaria tenuissima 469<br />

alternate host 610<br />

alternation of generations 155,<br />

160, 229<br />

alveoli 427<br />

Amanita 540–541; poisoning 540<br />

Amanita caesarea 540, Pl. 9<br />

Amanita fulva 540<br />

Amanita muscaria 526, 540–541, Pl. 9<br />

Amanita pantherina 540<br />

Amanita phalloides 540–541; poisoning<br />

540<br />

Amanita rubescens 521, 540; trama 524<br />

Amanita vaginata 492, 540<br />

Amanita verna 540<br />

Amanita virosa 540<br />

a-amanitin 539, 540<br />

Amani<strong>to</strong>psis 540<br />

Amauroascus 290<br />

Amauroderma rugosum 518


818 INDEX<br />

amerospore 23<br />

2-aminopyrimidine-type fungicides<br />

410<br />

amixis 176<br />

ammonia, as signal molecule 42<br />

amoebae 40, 41, 42, 49, 55; ‘social<br />

amoebae’ 44, 46<br />

amoeboid plasmodium 55, 60<br />

Ampelomyces quisqualis 412<br />

amphigynous fertilization 109<br />

amphithallism 506, 535, 538<br />

amphotericin B 260, 279, 291,<br />

306–307, 663<br />

amylase 282<br />

amyloid staining 28, 241; of ascus tip<br />

334; of ascus wall 414, 419; of<br />

basidiospore 566<br />

Amylostereum 571<br />

Anaeromyces 151<br />

anamorph 32, 38<br />

anamorph-teleomorph relationships<br />

32, 38, 299; see aero-aquatic fungi,<br />

aquatic hyphomycetes, Cordyceps,<br />

Epichloe, nema<strong>to</strong>phagous fungi<br />

Anaptychia ciliaris 448<br />

anas<strong>to</strong>mosis 3, 227, 377, 625<br />

anas<strong>to</strong>mosis groups 594<br />

andromorph 32<br />

anaerobic fungi 79, 150<br />

angiocarpic hymenophore 520, 521,<br />

522–525<br />

Anguillospora crassa (teleom. Mollisia<br />

uda) 685, 690<br />

Anguillospora furtiva (teleom. Pezoloma)<br />

690<br />

Anguillospora longissima (teleom.<br />

Massarina) 689, 692<br />

anisokont zoospore 24, 55, 64<br />

annellide 233, 234, 368<br />

annulus 520, 533<br />

anther smut 652<br />

antheridiol 88, 89, 90–91<br />

antheridium in Oomycota 29, 77, 85,<br />

89, 101; monoclinous and diclinous<br />

85, 101; in Ascomycota 296, 307, 416<br />

anthracnose 388, 389<br />

antifreeze proteins 537<br />

antioxidants 482, 670<br />

Aphanomyces 94–95<br />

Aphanomyces astaci 79, 94<br />

Aphanomyces euteiches 79, 94<br />

Aphanomyces patersonii 94<br />

aphanoplasmodium 48<br />

apical apparatus; see ascus<br />

apical paraphysis; see paraphysis<br />

Apiocrea chrysosperma; see Sepedonium<br />

chrysospermum<br />

aplanetism 92<br />

Aplanopsis 93<br />

aplanospore 2, 15, 24, 64, 165–214,<br />

216–223, 224, 225<br />

apomixis 85, 160<br />

apophysis 46; of basidiospore 490<br />

apop<strong>to</strong>sis in fungi 534<br />

apothecium 245, 246, Pl. 6, Pl. 7;<br />

cleis<strong>to</strong>hymenial 422;<br />

gymnohymenial 422;<br />

hemiangiocarpic 439<br />

apple, brown rot 306, 432, 433, 537;<br />

canker 343; cedar-apple rust 632;<br />

powdery mildew 390, 405; scab 478,<br />

479; spur canker 432<br />

appressorium, in Blumeria 396;<br />

in Botrytis 438; in Colle<strong>to</strong>trichum 388;<br />

in Magnaporthe 378–381; in<br />

Metarhizium 364; in Pip<strong>to</strong>cephalis 202;<br />

in Pythium, Phy<strong>to</strong>phthora 100; in<br />

Tapesia 440; in Uromyces 615, 616<br />

Aqualinderella fermentans 79<br />

aquatic fungi 577; see aero-aquatic<br />

fungi, aquatic hyphomycetes<br />

aquatic hyphomycetes 685, 686–696;<br />

adaptations <strong>to</strong> aquatic habitat 694;<br />

anamorph-teleomorph connections<br />

686; branched clamped conidia<br />

688–689; ecophysiology 694–696;<br />

foam 685, 694; invertebrate feeding<br />

695–696; sigmoid spores 686,<br />

689–691, 694; spore concentrations<br />

695; tetraradiate conidia 686,<br />

687–688, 693<br />

Arabidopsis thaliana 62, 118, 122<br />

arboriform hypha 518<br />

arbuscule (of VAM) 218, 220, 221<br />

archaeology 479, 610<br />

Archiascomycetes 250–260, Pl. 4<br />

Arcyria denudata 51, Pl. 1<br />

Armillaria 549–550; orchid mycorrhiza<br />

598; rhizomorphs 16, 17, 550<br />

Armillaria bulbosa 511, 550<br />

Armillaria cepistipes 550<br />

Armillaria lutea 550<br />

Armillaria luteobubalina 550<br />

Armillaria mellea 547, 549;<br />

dikaryotization (Buller<br />

phenomenon) 510; orchid<br />

mycorrhiza 550; unusual life cycle<br />

507<br />

Armillaria os<strong>to</strong>yae 550<br />

Armillaria tabescens 550<br />

Arrhenosphaera 286<br />

Arthrobotrys 676; induction of<br />

nema<strong>to</strong>de traps 682<br />

Arthrobotrys eudermata 677, 684;<br />

chlamydospores 684<br />

Arthrobotrys oligospora 676, 678<br />

Arthrobotrys robusta 678<br />

arthroconidium 281, 421, 422, 502,<br />

599, 600<br />

Arthroderma 293<br />

Articulospora tetracladia (teleom.<br />

Hymenoscyphus tetracladius) 685, 687,<br />

688<br />

Ascobolus 419–423; ascus development<br />

238; mating behaviour 420–421<br />

Ascobolus carbonarius 420<br />

Ascobolus crenulatus 420; and hyphal<br />

interference 540<br />

Ascobolus furfuraceus 420, 422, Pl. 6;<br />

apothecium development 422;<br />

hormonal control 421; mating 421<br />

Ascobolus immersus 420, 421; ascospore<br />

wall 237; gene recombination 423;<br />

mating 421; turgor pressure in<br />

ascus 242<br />

Ascobolus scatigenus, mating 420<br />

Ascobolus stic<strong>to</strong>ideus 422<br />

ascocarps (ascomata) 21, 245, 246;<br />

see epigeous, hypogeous<br />

Ascochyta (teleom. Didymella) 464<br />

Ascochyta fabae 464<br />

Ascochyta pisi 464, 465<br />

Ascochyta rabiei 464<br />

Ascocoryne 444<br />

ascogenous hypha 236, 296, 307, 309,<br />

318, 418<br />

ascogonium 296, 307, 309, 318, 416<br />

ascohymenial development 315, 459<br />

ascolocular development 455, 459<br />

ascomycetes, classification 247–249;<br />

phylogeny 248; significance 246–247<br />

ascorbic acid; see vitamin C<br />

Ascosphaera 286–287<br />

Ascosphaera apis 288, 289<br />

Ascosphaerales 286–289<br />

ascospore 25–26; alignment in ascus<br />

323–324; appendages 238, 322;<br />

cleavage 237, 267, 422; discharge<br />

242–245, 318; eclosion 336, 338;<br />

multi-spored projectiles 318, 420;<br />

puffing 243, 415, 427, 442; tendrils<br />

245; wall layers 321, 333


INDEX<br />

819<br />

ascospore-delimiting membrane<br />

(ascus vesicle) 237, 238, 422<br />

ascospore wall 237–239<br />

ascosporogenous yeast 265<br />

ascostroma 246, 459<br />

ascus apical apparatus 241, 242, 335,<br />

349, 358, 375; annulus 241;<br />

bilabiate 241; operculate versus<br />

inoperculate 241<br />

ascus development 236, 237<br />

ascus vesicle; see ascospore-delimiting<br />

membrane<br />

ascus wall 239, 240, 241; bitunicate<br />

239, 429, 459; ec<strong>to</strong>tunica 240, 459,<br />

460; endotunica 240, 459, 460;<br />

fissitunicate 239, 455, 459;<br />

pro<strong>to</strong>tunicate 241; rostrate 239,<br />

455; unitunicate 239<br />

Aseroe rubra 592, Pl. 11<br />

Ashbya; see Eremothecium<br />

aspergillomarasmine A 478<br />

aspergillosis 306<br />

Aspergillus 297–301; anamorphteleomorph<br />

relationships 308;<br />

cell-cycle 301; conidiophores 298,<br />

299; conidiogenesis 299–301;<br />

sclerotia 308; teleomorphs 299,<br />

308–310; uniseriate versus biseriate<br />

298<br />

Aspergillus carbonarius 305<br />

Aspergillus clavatus 306<br />

Aspergillus flavus 302, 306, 308;<br />

afla<strong>to</strong>xins 305<br />

Aspergillus fumigatus 306, 308<br />

Aspergillus nidulans (teleom. Emericella<br />

nidulans), conidiogenesis 299, 301;<br />

conidiophore 301, 302; parasexual<br />

cycle 230; penicillin production 303<br />

Aspergillus niger 308; citric acid<br />

production 303; conidiogenesis<br />

299<br />

Aspergillus ochraceus 305<br />

Aspergillus oryzae 276, 302, 305, 308<br />

Aspergillus parasiticus 302, 308, Pl. 4<br />

Aspergillus sojae 302<br />

Aspergillus versicolor 304<br />

astaxanthin 665, 671<br />

Asterionella formosa, and Rhizophydium<br />

141<br />

Asteromella 481<br />

asthma 306, 471, 483, 560, 647, 669;<br />

‘mushroom worker’s lung’ 542<br />

Astraeus 586; bole<strong>to</strong>id clade 579<br />

Atricordyceps 349, 681<br />

Aureobasidium pullulans 231, 261,<br />

484–485, 486, 668<br />

Auricularia auricula-judae 208, 493, 504,<br />

601, 602, Pl. 11; basidiospore<br />

germination 601; basidium 489, 602;<br />

dolipore septum 499; life cycle 603<br />

Auricularia mesenterica 602<br />

Auricularia polytricha, cultivation 602<br />

Auriculariales 601–604; basidial<br />

septation 601; lunate microconidia<br />

601, 602<br />

Auriscalpium 572<br />

Auriscalpium vulgare 567, 572; dolipore<br />

septum 499<br />

au<strong>to</strong>digestion 537<br />

au<strong>to</strong>ecious 613<br />

au<strong>to</strong>phagocy<strong>to</strong>sis 11, 13, 381<br />

au<strong>to</strong>phagosome 273<br />

Auxarthron 290<br />

auxiliary zoospore 76, 83<br />

auxins (of host plants) Albugo 122;<br />

Plasmodiophora 60; Taphrina 253;<br />

Ustilago maydis 644<br />

avirulence gene products 484, 619<br />

avoidance response (fugitropism) 168<br />

axoneme 68<br />

azoles 279, 291<br />

azygospore 179, 181<br />

bacterial endosymbionts of fungi:<br />

Geosiphon pyriforme 221; Laccaria 553;<br />

Morchella 428; Rhizopus microsporus<br />

183<br />

Baermann funnel 681<br />

ballis<strong>to</strong>spores 28; Ascomycota<br />

242–245; Basidiobolus 205;<br />

Basidiomycota 493, 495–496, 660;<br />

En<strong>to</strong>mophthora 216; Epicoccum 467;<br />

Erynia 213; Furia 219;<br />

Peronosclerospora 125<br />

banana, Panama wilt 347; postharvest<br />

anthracnose 388<br />

barberry and Puccinia graminis<br />

620–622; barberry eradication 622<br />

barley, covered smut 643; dwarf rust<br />

628; leaf-blotch 462; leaf-spot 478;<br />

loose smut 643; net-blotch 478;<br />

powdery mildew 398; stripe rust<br />

627; see mlo allele<br />

barrage phenomenon 325, 365, 428<br />

Basidiobolus microsporus 207<br />

Basidiobolus ranarum 203–208;<br />

germination patterns 204, 205–206;<br />

life cycle 208; pathogenicity 207;<br />

phylogeny 37, 166, 203; sexual<br />

reproduction 207, 209<br />

basidiocarps (basidiomata) 22,<br />

517–523; development 519–522,<br />

534–535, 545<br />

basidiomycete yeasts 658, 659,<br />

660–672; ecology 658–659;<br />

isolation 658; phylogeny 660, 661;<br />

versus ascomycete yeasts 659;<br />

see Heterobasidiomycetes,<br />

Urediniomycetes,<br />

Ustilaginomycetes<br />

Basidiomycota 487–513; asexual<br />

reproduction 501–506;<br />

classification 512; dikaryotization<br />

498; life cycle 492; mating systems<br />

506–510; mycelial cords 500;<br />

phylogeny 512; plasmogamy 498;<br />

relationship with Ascomycota 511;<br />

rhizomorphs 500; sclerotia 501;<br />

see basidium, basidiospore<br />

basidiospore 26, 27, 28; development<br />

490–492, 493; discharge 493–495;<br />

germination 496; numbers 495,<br />

497; see adaxial blob, Buller’s drop,<br />

hilar appendix, hilar appendix<br />

body, surface tension catapult<br />

basidium; development 488–490;<br />

morphology 487–488; nuclear<br />

events 490, 492; holobasidium<br />

versus phragmobasidium 488;<br />

probasidium versus metabasidium<br />

490; chias<strong>to</strong>basidium versus<br />

stichobasidium 490;<br />

see epibasidium, furcate basidium,<br />

heterobasidium, phragmobasidium,<br />

tremelloid basidium<br />

bassianolide 363<br />

Batkoa 203<br />

Beaumont period 114<br />

Beauveria bassiana 361–362, 363; in<br />

biological control 363<br />

beauvericin 363<br />

bellows mechanism 580–581, 588, 639<br />

benomyl 410; resistance 410<br />

benzaldehyde 630<br />

benzimidazole fungicides 410<br />

benzothiadiazole 410, 412<br />

Bettsia 286<br />

Beverwijkella pulmonaria 697, 700,<br />

701<br />

bifac<strong>to</strong>rial mating 507<br />

binding hypha 518<br />

biological clock 331


820 INDEX<br />

biological control 263; of fungi 340,<br />

342, 347, 377, 412–413, 433–434,<br />

485, 551, 569, 596, 668; of insects<br />

211, 363–364; of nema<strong>to</strong>des 684; of<br />

weeds 631<br />

bioremediation 531<br />

biotechnology; see bread-making,<br />

brewing, biological control,<br />

biotransformation, citric acid<br />

fermentation, enzyme production,<br />

food production, griseofulvin,<br />

penicillin, riboflavin production,<br />

saké, wine production<br />

biotransformation 167, 197<br />

biotrophy in mycoparasites 191, 201;<br />

in plant pathogens; see Blumeria<br />

graminis, Cladosporium fulvum,<br />

Exobasidiales, Peronosporales,<br />

Uredinales, ustilaginomyce<strong>to</strong>us<br />

smuts<br />

bipolar; see heterothallism<br />

Bipolaris 471–472<br />

Bipolaris sorokiniana (teleom.<br />

Cochliobolus sativus) 472, 474<br />

birch, rust 635; sap and<br />

Xanthophyllomyces 665; witches’<br />

broom 251, Pl. 4<br />

bird’s nest fungi; see Crucibulum cyathus<br />

Biscogniauxia mediterranea 333<br />

Biscogniauxia nummularia 333<br />

biseriate; see Aspergillus<br />

Bisporella citrina 444, Pl. 7<br />

bitunicate; see ascus wall<br />

biverticillate; see Penicillium<br />

conidiophores<br />

Bjerkandera adusta 519, 562<br />

black yeasts 486<br />

Blakeslea trispora 174, 194, 195<br />

Blakeslea unispora 195<br />

blastic conidiogenesis 231;<br />

see holoblastic, enteroblastic<br />

Blas<strong>to</strong>cladiales 153–162; bipolar<br />

germination 155; life cycles<br />

155–160; zoospore 153–155<br />

Blas<strong>to</strong>cladiella 160–162; Cys<strong>to</strong>cladiella<br />

160; Eucladiella 160<br />

Blas<strong>to</strong>cladiella emersonii 153, 154,<br />

161, 162<br />

Blas<strong>to</strong>cladiella variabilis 160<br />

blas<strong>to</strong>conidium 435, 484, 503, 504<br />

Blas<strong>to</strong>myces dermatitidis 289, 290–292<br />

Blumeria graminis 390, 393, 394,<br />

395–399, 400, 401; appressorium<br />

396; attachment 395; chasmothecia<br />

400, 401; conidia 395, 399; formae<br />

speciales 395; gene-for-gene<br />

interactions 398; haus<strong>to</strong>rium 393,<br />

394, 397–398, 399; infection process<br />

395, 396, 397; life cycle 399–401;<br />

primary germ tube 396; races 395,<br />

408; secondary germ tube 396;<br />

signalling 396<br />

‘boat-hook’ spines 82, 84<br />

bole<strong>to</strong>id clade 555–560, Pl. 9;<br />

gasteromycetes 579, 585–588;<br />

phylogeny 516–517, 555, 585<br />

Boletus 557–558; related <strong>to</strong><br />

gasteromycetes 578<br />

Boletus appendiculatus 557<br />

Boletus badius 515, 557, Pl. 9<br />

Boletus chrysenteron 556, 557<br />

Boletus edulis 556, 557; cultivation 526,<br />

557; trama 524<br />

Boletus erythropus 557, Pl. 9<br />

Boletus luridus 557<br />

Boletus parasiticus 557<br />

Boletus rubinellus 489; basidiospore<br />

development 491–492, 493<br />

Boletus satanas 557<br />

boom-and-bust cycle 619<br />

Bordeaux mixture 112, 119, 253,<br />

410, 647<br />

bothrosome; see sagenogen<br />

Botryotinia fuckeliana 430, 434–435;<br />

see Botrytis cinerea<br />

Botryotinia porri 436<br />

Botryotrichum (teleom. Chae<strong>to</strong>mium) 332<br />

Botrytis cinerea (teleom. Botryotinia<br />

fuckeliana) 4, 7, 32, 121, 403,<br />

434–435; attachment 438; fungicide<br />

resistance 439; life cycle 437;<br />

macroconidia 434, 436;<br />

microconidia 435; pathogenicity<br />

mechanisms 436–439, 475; sclerotia<br />

19; septum 228<br />

Bovista 581<br />

bow-tie reaction 570, 571<br />

bramble rust 631<br />

brassicas, Albugo white blisters 122,<br />

125; blackleg 463; club-root 55–56,<br />

57; powdery mildew 402; see also<br />

oilseed rape<br />

bread-making 274<br />

breeding for resistance (in plants) 63,<br />

119, 138, 347, 368, 408–410, 440,<br />

625–627, 633–634, 642<br />

Bremia lactucae 120, 121, 122<br />

Brettanomyces 275<br />

brewing 274–275<br />

broad bean, anthracnose 387, Pl. 5;<br />

leaf- and pod-spot 464; net-blotch<br />

466; rust 631<br />

brown-rot (of wood) 529, 529–530,<br />

556, 559–560, 563–564<br />

bubble-trap propagules 696, 698<br />

bud scar 271, 283<br />

bulbil 490, 505<br />

Bulbillomyces farinosus (anam. Aegerita<br />

candida) 505, 506, 697<br />

bulbous hyphae 378, 381; see infection<br />

hyphae<br />

Bulgaria inquinans 445, Pl. 7<br />

Buller phenomenon 510, 570<br />

Buller’s drop 493–494, 495<br />

Bullera 659<br />

Bulleromyces (anam. Bullera) 659<br />

bunt 637; common bunt 647; dwarf<br />

bunt 649; Karnal bunt 649<br />

Byssochlamys 307–308<br />

Byssochlamys nivea 306, 307<br />

Ca 2þ channels, stretch-activated 9–10,<br />

66, 615<br />

Cadophora 439<br />

Caecomyces 151<br />

caeoma 635<br />

Caespi<strong>to</strong>theca forestalis 393<br />

calcineurin signalling 664<br />

calciseptine 646<br />

callose 397<br />

calmodulin 384, 385<br />

Calocera cornea 598, 601<br />

Calocera viscosa 600, 601, Pl. 11;<br />

basidiospore germination 600, 601;<br />

basidium 489<br />

Calos<strong>to</strong>ma 585<br />

Calvatia 581<br />

Calvatia excipuliformis 582, Pl. 11<br />

Calvatia gigantea 581, 582<br />

cAMP (cyclic AMP) 41, 105, 384, 645; as<br />

hormone 42<br />

Candelabrum desmidiaceum 701<br />

Candida 276–280<br />

Candida albicans 226, 276–280, 291, 654;<br />

candidiasis 278–280; cryptic sexual<br />

phase 277; dimorphism 277; drug<br />

resistance 279–280;mating 277–278;<br />

switching 278; treatment 279–280<br />

Candida dubliniensis 276, 280<br />

Candida glabrata 276, 278, 280<br />

Candida inconspicua 276<br />

Candida krusei 276, 280


INDEX<br />

821<br />

Candida parapsilosis 227, 280<br />

Candida tropicalis 280<br />

Candida utilis; see Pichia jadinii<br />

candidiasis 278–280; treatment<br />

279–280<br />

canker (of plants) 333, 342, 343,<br />

376, 634; self-healing 376<br />

cantharelloid clade 574–575;<br />

phylogeny 516–517<br />

Cantharellus 574<br />

Cantharellus cibarius 574, Pl. 10;<br />

cultivation 526; post-meiotic<br />

mi<strong>to</strong>sis 490<br />

Cantharellus cinnabarinus 574<br />

Cantharellus tubaeformis 574<br />

canthaxanthin 574<br />

capilliconidia 205, 206<br />

capillitium 51, 581<br />

Capronia 486<br />

captan 479, 651<br />

carbendazim 410<br />

carboxin 651<br />

b-carotene 170, 175, 195, 434, 665,<br />

Pl. 12<br />

carotenoids 143, 150, 156, 185, 253,<br />

414, 456, 574, 598, 605, 613<br />

caspofungin 279, 291<br />

Catenaria allomycis 153<br />

Catenaria anguillulae 153, 681<br />

cattle; see grazing animals<br />

Caulochytrium 127<br />

caulocystidium 525<br />

cdc2 51, 256<br />

cell cycle 51, 256–257, 270–272<br />

cellulase 339–340, 396, 530<br />

cellulose 528; in fungal cell walls 6,<br />

42, 67, 75, 127; as substrate 148,<br />

150, 529–530<br />

cell wall 5, 6; in Mucorales 167; in<br />

Saccharomyces cerevisiae 270<br />

cell wall, biosynthesis 6–8<br />

cell wall-degrading enzymes 375, 396,<br />

438, 596<br />

centrum 242, 318<br />

cephalodium (in lichens) 451<br />

cephalosporins 303–304, 350<br />

Cephalotrichum stemonitis 230, 368, 370<br />

Ceratiomyxa fruticulosa 46, 47, Pl. 1<br />

Cera<strong>to</strong>basidiales 594–598;<br />

basidiocarps 595; basidium 595;<br />

see Rhizoc<strong>to</strong>nia<br />

Cera<strong>to</strong>basidium 594; see Rhizoc<strong>to</strong>nia<br />

Cera<strong>to</strong>cystis 369–373; versus Ophios<strong>to</strong>ma<br />

364, 373<br />

Cera<strong>to</strong>cystis coerulescens 373<br />

Cera<strong>to</strong>cystis fagacearum 369, 373<br />

Cera<strong>to</strong>cystis fimbriata 369<br />

Cera<strong>to</strong>rhiza (teleom. Cera<strong>to</strong>basidium) 594<br />

cera<strong>to</strong>-ulmin 366, 368<br />

Cercospora 481–482<br />

Cercospora beticola 481, 484<br />

Cercospora coffeicola 481<br />

Cercospora zea-maydis 481<br />

cercosporin 482, 485<br />

cereal rusts 627–629; origin 628<br />

cereals, downy mildews 126; ergot<br />

352; eyespot 439, 441; leaf blotch<br />

439; mosaic viruses 62; powdery<br />

mildew 393–399, 400, 401; sharp<br />

eyespot 596; take-all 385–386; see<br />

also individual cereals<br />

Chaenothecopsis 246<br />

Chae<strong>to</strong>cladium brefeldii 191, 192<br />

Chae<strong>to</strong>cladium jonesii 191<br />

Chae<strong>to</strong>mium 331–332; perithecial<br />

development 332<br />

Chae<strong>to</strong>mium brasiliense 332<br />

Chae<strong>to</strong>mium cochliodes 331<br />

Chae<strong>to</strong>mium elatum 331–332<br />

Chae<strong>to</strong>mium globosum 332<br />

Chae<strong>to</strong>mium piluliferum 332<br />

Chae<strong>to</strong>mium thermophile 331<br />

Chae<strong>to</strong>thyriales 486<br />

chalk brood disease; see honey bees<br />

chasmothecium 246, 392, 401,<br />

409; in oversummering 401;<br />

viability 405<br />

cheese production 303<br />

cheilocystidium 525<br />

chemotaxis 42, 100, 106<br />

chemotaxonomy 454<br />

chemotropism 201, 330, 586<br />

Chernobyl accident 454<br />

cherry, witches’ broom 251; silver-leaf<br />

571<br />

chestnut, blight 376<br />

chias<strong>to</strong>basidium 490<br />

chickpea, blight 464<br />

chitin (in cell walls) 5, 6, 55, 67, 127,<br />

153, 165, 255, 538; bud scar 261;<br />

spiral growth 171<br />

chi<strong>to</strong>san (in cell walls) 6, 165<br />

chi<strong>to</strong>somes 6, 7, 168<br />

chlamydospore 29, 30, 81, 105, 218,<br />

345, 347, 484, 486, 504, 505, 545,<br />

667; see teliospore<br />

3-chloroanisylalcohol 530, 531, 532<br />

Chlorociboria 444<br />

Chlorociboria aeruginascens Pl. 7<br />

chloromethane 532<br />

chlorothalonil 631<br />

Choanephora cucurbitarum 173, 179,<br />

193, 194, 202<br />

Chondrostereum purpureum 571, Pl. 10<br />

Chromista 67; see Straminipila<br />

Chrysonilia 327, 329<br />

Chrysosporium 290, 293, 294<br />

Chytridiales 134–145<br />

Chytridiomycota 71, 127–164, Pl. 3;<br />

classification 134; flagellar<br />

apparatus 129, 130; germination<br />

(monopolar versus bipolar) 131;<br />

inoperculate and operculate species<br />

129; life cycles 131; microbody-lipid<br />

complex 130, 131; sexual<br />

reproduction 132–133; thallus 128;<br />

zoospore 129–131<br />

Chytriomyces hyalinus 132<br />

cinerean 438<br />

circadian rhythm 331, 590<br />

cirrhus 315, 332, 465<br />

citric acid fermentation 303<br />

citrinin 305<br />

citrus fruit rot Pl. 4; see Penicillium<br />

italicum, P. digitatum<br />

Cladina 458<br />

Cladina rangiferina 458, Pl. 8<br />

Cladina stellaris 458<br />

Cladochytrium 142–144<br />

Cladochytrium replicatum 142, 143, 144<br />

Cladonia 448, 457–458<br />

Cladonia convoluta 452<br />

Cladonia cristatella 450<br />

Cladonia floerkiana 457, Pl. 8<br />

Cladonia pyxidata 450, 457<br />

Cladophialophora 486<br />

Cladosporium 481–484; Mycosphaerella<br />

teleomorphs 482; spores 22–23<br />

Cladosporium cladosporioides 483<br />

Cladosporium echinulatum 482<br />

Cladosporium fulvum 483–484, 619;<br />

biotrophy 483<br />

Cladosporium herbarum 482–483, 485<br />

Cladosporium humile 482<br />

Cladosporium macrocarpum 485<br />

clamp connection 499, 500–501, 569;<br />

versus crozier 512<br />

Clas<strong>to</strong>derma 51<br />

Clathrosphaerina zalewskii (teleom.<br />

Hyaloscypha) 697, 699–700<br />

Clathrus archeri 591–592<br />

Clathrus ruber 591–592, Pl. 11


822 INDEX<br />

Clavaria 575–576; ericoid mycorrhiza<br />

576<br />

Clavaria argillacea 576<br />

Clavaria vermicularis 576<br />

Clavariadelphus pistillaris 515, 576<br />

Clavariopsis aquatica (teleom.<br />

Massarina) 688<br />

Clava<strong>to</strong>spora longibrachiata 685<br />

Clava<strong>to</strong>spora stellata 685<br />

Claviceps 349<br />

Claviceps africana 349, 355<br />

Claviceps fusiformis 349, 355<br />

Claviceps microcephala 350<br />

Claviceps paspali 349, 355<br />

Claviceps phalaridis 355<br />

Claviceps purpurea 349–351, 352,<br />

353–355; alkaloids 353, 354, 355;<br />

ascospore discharge 241; control<br />

355; ergotism 352–353; life cycle<br />

349–350, 351, 352; sclerotia 19, 20,<br />

350, 352<br />

Claviceps purpurea var. spartinae 355,<br />

Pl. 5<br />

Claviceps sorghi 349<br />

Clavicipitales 316, 348–364; alkaloids<br />

349<br />

Clavicorona pyxidata 489<br />

Clavulina 575–576<br />

Clavulina cristata 574, 575<br />

Clavulina rugosa 575<br />

cleavage of cy<strong>to</strong>plasm 98, 106,<br />

158, 172<br />

cleis<strong>to</strong>hymenial; see apothecium<br />

cleis<strong>to</strong>thecium 245, 246, 285, 309,<br />

313, 314<br />

clep<strong>to</strong>biosis 451, 456<br />

Cli<strong>to</strong>cybe 552<br />

Cli<strong>to</strong>cybe clavipes 521<br />

Cli<strong>to</strong>cybe geotropa 552<br />

Cli<strong>to</strong>cybe nebularis 552, 553<br />

Cli<strong>to</strong>cybe odora 552<br />

club root 55–56, 57; see Plasmodiophora<br />

brassicae<br />

CO 2 fixation 162<br />

Coccidioides immitis 290–293<br />

Coccidioides posadasii 293<br />

Coccomyxa 452, 456<br />

Cochliobolus 460, 471–477; Bipolaris<br />

anamorphs 475; mating type<br />

idiomorphs 477; pathology 475–477<br />

Cochliobolus carbonum 477; HC-<strong>to</strong>xin<br />

476<br />

Cochliobolus cymbopogonis, ascus<br />

discharge 243, 244<br />

Cochliobolus heterostrophus (anam.<br />

Bipolaris maydis) 477; T-<strong>to</strong>xin 476<br />

Cochliobolus sativus (anam. Bipolaris<br />

sorokiniana) 475–476;<br />

prehelminthosporol 476<br />

Cochliobolus vic<strong>to</strong>riae 476–477, 619;<br />

origin 477; vic<strong>to</strong>rin C 476<br />

cocoa, pod rot 103; frosty pod rot 551;<br />

witches’ broom 547, 550<br />

Coelomomyces 127, 153<br />

coelomycetes 231<br />

coenocytic hypha 2, 75, 167<br />

coffee, berry disease 388; eyespot 481;<br />

leaf rust 632–634<br />

collagen 654<br />

Colle<strong>to</strong>trichum 231, 387, 389; infection<br />

strategies 389<br />

Colle<strong>to</strong>trichum capsici 388<br />

Colle<strong>to</strong>trichum coccodes 388–389<br />

Colle<strong>to</strong>trichum coffeanum 388<br />

Colle<strong>to</strong>trichum destructivum 388<br />

Colle<strong>to</strong>trichum gloeosporioides 387, 389<br />

Colle<strong>to</strong>trichum gossypii 388<br />

Colle<strong>to</strong>trichum graminicola 387, 388<br />

Colle<strong>to</strong>trichum lindemuthianum 388,<br />

Pl. 5<br />

Colle<strong>to</strong>trichum musae 388–389<br />

Collybia 546, 552<br />

Collybia butyracea post-meiotic mi<strong>to</strong>sis<br />

490<br />

Collybia tuberosa 501<br />

columella; in puffballs 584; in<br />

Zygomycota 166, 167, 182, 184–185,<br />

192, 208<br />

Comatricha 51<br />

co-metabolism 531<br />

compound appressorium 595<br />

Conidiobolus 203, 208–211; eversion<br />

mechanism in conidium discharge<br />

208<br />

Conidiobolus coronatus 209, 210, 211;<br />

insecticidal <strong>to</strong>xin 211; loriconidia<br />

211; villose conidia 210<br />

conidiogenesis 31, 231–235, 299,<br />

300, 301; blastic versus thallic 232;<br />

hormonal control 301; see holoblastic,<br />

enteroblastic, thallic<br />

conidiogenone 301<br />

conidioma 230<br />

conidiophore 230, 298, 310<br />

conidium (conidiospore) 30–32;<br />

unitunicate versus bitunicate<br />

(En<strong>to</strong>mophthorales) 203, 215;<br />

see arthroconidium,<br />

blas<strong>to</strong>conidium, oidium,<br />

phialoconidium, spermatium<br />

conifers, butt rot 567; root rot 573,<br />

Pl. 2<br />

Coniophora 558<br />

Coniophora puteana 570<br />

Coniothyrium minitans 433<br />

conjugate nuclear division 499<br />

connective 300, 327<br />

convergent evolution 673, 686,<br />

693, 701<br />

copper, as fungicide 479, 633, 647,<br />

651; see Bordeaux mixture<br />

coprine 538, 539<br />

Coprinellus 537<br />

Coprinopsis 537<br />

Coprinus 536–540; life cycle 492;<br />

mating behaviour 538; sclerotium<br />

537; stipe elongation 538<br />

Coprinus atramentarius 522, 537–538;<br />

and alcohol 538<br />

Coprinus cinereus 22, 489, 498, 501, 521,<br />

533, 538; basidiospore development<br />

491, 492, 493, 494; dikaryotization<br />

(Buller phenomenon) 510;<br />

gravitropism 549; mating type<br />

alleles 508, 509; oidia 502, 503<br />

Coprinus comatus 533, 537; mating<br />

system 506; phylogeny 536; related<br />

<strong>to</strong> gasteromycetes 578<br />

Coprinus congregatus 499<br />

Coprinus curtus 537<br />

Coprinus domesticus 537<br />

Coprinus ephemerus f. bisporus, mating<br />

system 506<br />

Coprinus heptemerus, hyphal<br />

interference 540<br />

Coprinus micaceus 537<br />

Coprinus plicatilis 537<br />

Coprinus psychromorbidus 537<br />

Coprinus sterquilinus, mating system<br />

506; phylogeny 536; sclerotia<br />

537–538<br />

Coprobia granulata 229<br />

coprophilous fungi 238; Ascobolus 419;<br />

Basidiobolus 203–208; Coprinus 537;<br />

Mucor 181; Pilaira 189–190; Pilobolus<br />

185–189<br />

Cordyceps 360–362; perithecial stroma<br />

362; sclerotium 360<br />

Cordyceps bassiana (anam. Beauveria<br />

bassiana) 362<br />

Cordyceps brittlebankisoides (anam.<br />

Metarhizium anisopliae) 363


INDEX<br />

823<br />

Cordyceps capitata 362<br />

Cordyceps militaris 20, 361, Pl. 5<br />

Cordyceps ophioglossoides 314, 362, Pl. 4;<br />

host-jumping 362<br />

Cordyceps sinensis 361, 362<br />

Cordyceps subsessilis (anam.<br />

Tolypocladium inflatum) 349, 364<br />

coremium 230, 310, 311<br />

Coriolus; see Trametes<br />

corn; see maize<br />

correlated rust species 614<br />

cortina 520, 553, 555<br />

Cortinarius 555; related <strong>to</strong><br />

gasteromycetes 578<br />

Cortinarius orellanus 555<br />

Cortinarius purpureus 553<br />

Costantinella cristata 428<br />

cot<strong>to</strong>n, anthracnose 388<br />

Craterellus cornucopioides 574<br />

crayfish plague (Aphanomyces astaci)<br />

94<br />

Cribraria 51<br />

Crinipellis 546<br />

Crinipellis perniciosa 547, 550<br />

Crinipellis roreri 551<br />

Cronartium 634–635; teliospore 634<br />

Cronartium flaccidum 634<br />

Cronartium quercuum 634<br />

Cronartium ribicola 634; canker 634<br />

crook root; see watercress<br />

crozier 236, 418; versus clamp<br />

connection 512<br />

Crucibulum 582, 585<br />

Crucibulum vulgare, mating-type<br />

fac<strong>to</strong>rs 509<br />

cruciform mi<strong>to</strong>sis; see Plasmodiophora<br />

brassicae<br />

crus<strong>to</strong>se lichens 447, Pl. 8<br />

Cryphonectria parasitica 375–377;<br />

basidium 662; biological control<br />

377; hypovirulence 376–377;<br />

Hypovirus 376–377; mating types<br />

662; pheromones 377; vegetative<br />

incompatibility 377; virus<br />

transmission 377<br />

Cryp<strong>to</strong>coccus 661<br />

Cryp<strong>to</strong>coccus neoformans 608,<br />

661–665; infection biology 663–664;<br />

life cycle 662, 663; mucilage<br />

capsule 661, 664; natural habitat<br />

662; switching of phenotypes 664;<br />

var. gattii 661; var. grubii 661; var.<br />

neoformans 661; virulence fac<strong>to</strong>rs<br />

664–665<br />

Ctenomyces serratus 293, 294<br />

cucumber necrosis virus (CNV) 148<br />

cultivation of mushrooms 525,<br />

602, 607; mycorrhizal fungi 527,<br />

558, 566; see Agaricus bisporus,<br />

Auricularia polytricha, Lentinula<br />

edodes, Tremella fuciformis<br />

Cunninghamella 196<br />

Cunninghamella bertholettiae 196<br />

Cunninghamella echinulata 196<br />

Cunninghamella elegans 196<br />

Curvularia 471–472<br />

Curvularia cymbopogonis 472, 475<br />

Curvularia lunata (teleom. Cochliobolus<br />

lunatus) 475<br />

cutinase 396, 438, 615<br />

Cyathus 582–585; splash cup 582<br />

Cyathus olla 582<br />

Cyathus stercoreus 582, 584<br />

Cyathus striatus 582; mating type<br />

fac<strong>to</strong>rs 509<br />

cyazofamid 113<br />

cycloheximide 373, 634<br />

cyclosporin A 350, 364, 664<br />

Cylindrocarpon 342, 345, 348<br />

Cylindrocarpon destructans (teleom.<br />

Nectria radicicola) 348<br />

Cylindrocarpon heteronema (teleom.<br />

Nectria galligena) 348<br />

Cylindrosporium concentricum<br />

(teleom. Pyrenopeziza brassicae) 439<br />

cystesium 525<br />

cystidiole 525<br />

cystidium 522, 524, 525<br />

cy<strong>to</strong>chalasins 364<br />

cy<strong>to</strong>kinins (of host plants), in<br />

Plasmodiophora 60; in Taphrina 253<br />

cy<strong>to</strong>skele<strong>to</strong>n 8, 9–10, 81, 258, 646; in<br />

spore cleavage 87, 158; see actin,<br />

microtubules<br />

Cyttaria 445<br />

Cyttaria darwinii Pl. 7<br />

Dacrymycetales 598–601; epibasidia<br />

599–600; life cycle 598<br />

Dacrymyces stillatus 493, 598–599, 600;<br />

basidiospore germination 599, 600<br />

Dacrymyces unisporus 598<br />

Dactylelina 675–676; induction of<br />

nema<strong>to</strong>de traps 683<br />

Dactylelina hap<strong>to</strong>tyla 679<br />

Daldinia 334<br />

Daldinia concentrica 236, 333–335, Pl. 5<br />

Daldinia fissa 334<br />

Daldinia loculata 334<br />

damping-off 96, 97, 102, 595, Pl. 2<br />

dandruff and Malassezia 672<br />

Dasyscyphus 25<br />

Dasyscyphus virgineus Pl. 7<br />

trans-2-decenoic acid, nematicidal<br />

properties 683<br />

decurrent gills 523<br />

de-dikaryotization 502, 630, 653<br />

de-diversification 661<br />

defence mechanisms; see host defence<br />

demethylation inhibi<strong>to</strong>rs (DMI) 651;<br />

see imidazoles, triazole<br />

demicyclic 613<br />

Dendrospora erecta 688, 690<br />

Dendrosporomyces prolifer 689<br />

Dendrosporomyces splendens 689<br />

dense-body vesicles 87, 107; see<br />

fingerprint vacuoles<br />

depletion zone 222<br />

Dermateaceae 439–440<br />

dermatitis and Malassezia 672<br />

derma<strong>to</strong>phytes 293<br />

desert truffle 427<br />

destruxins 364<br />

dextrinoid staining (with iodine) 241<br />

2,4-diacetylphloroglucinol 386<br />

diacylglycerol (DAG) 384<br />

diallyl disulphide 434, 435<br />

Diaporthales 316, 373–377<br />

Diaporthe (anam. Phomopsis) 373–375,<br />

605; ecology 375<br />

Diaporthe ambigua 375<br />

Diaporthe helianthi 375<br />

Diaporthe phaseolorum 373–375;<br />

pycnidium 374; sexual<br />

reproduction 374<br />

Diaporthe <strong>to</strong>xica 375<br />

diazonium blue B stain 659<br />

Dicho<strong>to</strong>mocladium 191<br />

diclinous antheridium 85<br />

Dictydium 51<br />

Dictyophora; see Phallus<br />

dictyospore 23, 469, 470<br />

Dictyosteliomycetes (dicy<strong>to</strong>stelid<br />

cellular slime moulds) 41–45<br />

Dictyostelium discoideum 42–45;<br />

aggregation 42, 44–45; cheater<br />

strains 44; life cycle 43<br />

Dictyuchus sterile 91, 93–94<br />

Didymella (anam. Ascochyta) 464<br />

didymospore 23<br />

diffluent sporangium 173<br />

diffuse extension growth 538


824 INDEX<br />

Digita<strong>to</strong>spora marina 693<br />

dikaryotization 498, 613; legitimate<br />

and illegitimate 510; see Buller<br />

phenomenon<br />

dimethyl disulphide 436, 591<br />

dimethyl sulphide 424<br />

dimitic basidiocarp construction 519<br />

dimorphism, sexual 436; yeast-hypha<br />

3, 181, 226; in Candida albicans 227,<br />

277; in Onygenaceae 291; in<br />

phialidic conidiogenesis 301, 302;<br />

in smut fungi 638<br />

Dioszegia 659<br />

diplanetism 81, 85<br />

diploid Eumycota 277, 666–667<br />

Dipodascus 282<br />

Dirina massiliensis f. sorediata 447,<br />

454<br />

Discosphaerina fulvida (anam.<br />

Aureobasidium pullulans) 484<br />

Discula destructiva 373<br />

disease forecasting, Hemileia vastatrix<br />

633; Phy<strong>to</strong>phthora infestans 113;<br />

Uncinula neca<strong>to</strong>r 403; Venturia<br />

inaequalis 479<br />

dispersal; see insects; mites<br />

dissepiments 519<br />

dis<strong>to</strong>septate 472, 473–474, 476, 478<br />

dithiocarbamates 112, 119, 253, 410<br />

dogwood, anthracnose 373<br />

dolipore septum 497, 498, 499, 594,<br />

597, 606; breakdown for nuclear<br />

passage 497, 600<br />

Dora<strong>to</strong>myces; see Cephalotrichum<br />

dormant kine<strong>to</strong>some;<br />

see kine<strong>to</strong>some<br />

Dothideales 480–486; lichens 455<br />

double discharge of asci 405<br />

downy mildews Pl. 2;<br />

see Peronosporales<br />

Drechmeria coniospora 349, 681, 682;<br />

adhesion <strong>to</strong> nema<strong>to</strong>des 683;<br />

infection process 684<br />

Drechslera (teleom. Pyrenophora) 471,<br />

477–478<br />

Drechslera tritici-repentis 476<br />

Drechslerella 677, 680; induction of<br />

nema<strong>to</strong>de traps 683; ring<br />

constriction 678<br />

Drechslerella brochopaga 679<br />

Drechslerella dactyloides 679<br />

drugs (against human pathogens) 279,<br />

280; see allyl amines,<br />

amphotericin B, azoles,<br />

caspofunginterbinafine,<br />

5’-fluorocy<strong>to</strong>sinegriseofulvin,<br />

griseofulvin, terbinafine, triazole<br />

dry rot 558–560; control 560<br />

Dudding<strong>to</strong>nia flagrans; see Arthrobotrys<br />

eudermata<br />

Dumontinia; see Sclerotinia tuberosa<br />

Dutch elm disease 366–368, Pl. 5;<br />

control 368<br />

earth stars 588<br />

Echinobotryum 368, 370<br />

echinocandins 260, 280<br />

Echinosteliales (Myxomycota) 51<br />

Echinostelium 51<br />

eclosion; see ascospore<br />

ec<strong>to</strong>mycorrhiza (sheathing<br />

mycorrhiza) 21, 314, 424, 526, 527,<br />

552, 573, 585–586<br />

ec<strong>to</strong>tunica; see ascus wall<br />

efrapeptins 364<br />

eelgrass (Zostera) wasting 72<br />

Elaphomyces 313–314, Pl. 4<br />

Elaphomyces granulatus 313<br />

Elaphomyces muricatus 313<br />

elater 51, 52<br />

electric field (around hypha) 13, 14<br />

elm, Dutch elm disease 366–368<br />

Emericella (anam. Aspergillus) nidulans<br />

229, 310; Woronin bodies 228;<br />

see Aspergillus nidulans<br />

encystment (of zoospores) 68, 76–77,<br />

83, 84, 87, 100, 106, 131<br />

endocy<strong>to</strong>sis 11, 13, 272<br />

Endogone 221<br />

Endogone flammicorona 221<br />

Endogone lactiflua Pl. 3<br />

endophytic fungi 333–334, 429, 484;<br />

of grasses see Neotyphodium; related<br />

<strong>to</strong> pathogens 347, 359, 375, 389,<br />

441, 462; xylotropic endophyte 336<br />

endoplasmic reticulum (ER) 11<br />

endospore, endosporium (of spore<br />

walls) 26, 417<br />

endospores 205, 206<br />

endotunica; see ascus wall<br />

enteroblastic conidiogenesis 31, 232,<br />

299, 300, 659<br />

en<strong>to</strong>mopathogenic fungi; see insect<br />

pathogens<br />

En<strong>to</strong>mophaga 203<br />

En<strong>to</strong>mophthora 203<br />

En<strong>to</strong>mophthora muscae 215–217, Pl. 3;<br />

infection 216, 218; penetration 216<br />

En<strong>to</strong>mophthora sepulchralis 217, 218<br />

En<strong>to</strong>mophthorales 202–217<br />

enzyme production 303, 339–340,<br />

562<br />

epibasidium 593, 595<br />

Epichloe 355–357; ascospore<br />

germination 356, 358; infection of<br />

grasses 356; spermatization by<br />

insects 356<br />

Epichloe baconii 355, 357<br />

Epichloe clarkii 356<br />

Epichloe festucae 356, 359<br />

Epichloe typhina 355, 358, Pl. 5<br />

Epicoccum 30<br />

Epicoccum nigrum 466; active conidial<br />

discharge 466<br />

Epidermophy<strong>to</strong>n 293<br />

epigeous fruit-bodies 414<br />

epiphragm 583, 584<br />

epispore, episporium (of spore walls)<br />

26, 417<br />

Eremascus 286–287<br />

Eremascus albus 287<br />

Eremascus fertilis 287<br />

Eremothecium 262, 284<br />

Eremothecium ashbyi 284<br />

Eremothecium coryli 284, 304<br />

Eremothecium gossypii 284<br />

ergometrine 353, 354<br />

ergosterol 279; biosynthesis 279, 412<br />

ergot 350, 352<br />

ergotamine 353, 354<br />

ergotism 352–353<br />

ergovaline 360<br />

ericoid mycorrhiza 444, 576<br />

Erynia 203, 211–215; resting bodies<br />

211<br />

Erynia conica 213–215; conidial types<br />

213, 214<br />

Erynia neoaphidis 211, 212, 213;<br />

asexual reproduction 213; infection<br />

of aphids 212–213; penetration 212;<br />

pro<strong>to</strong>plasts 212<br />

Erysiphales 390–413; appendages 391,<br />

392–393; chasmothecium 392;<br />

conidia 390; conidial surface and<br />

taxonomy 394, 395; control<br />

408–413; phylogeny and evolution<br />

392–393<br />

Erysiphe 401, 402, 403–404<br />

Erysiphe betae 402<br />

Erysiphe cruciferarum 401<br />

Erysiphe heraclei 402<br />

Erysiphe pisi 398, 402


INDEX<br />

825<br />

Erysiphe polygoni 402<br />

Erysiphe trifolii 402; resistance<br />

breeding 408<br />

ethanol; see alcoholic fermentation<br />

ethirimol 410<br />

ethnomycology 541<br />

ethylene 389<br />

euagarics clade 516–517, 532–555,<br />

Pl. 9; gasteromycetes 579, 581–585<br />

Eumycota 38<br />

Eupenicillium 311–313<br />

Eupenicillium crus<strong>to</strong>sum 313<br />

Eurotiales 297–314; anamorphteleomorph<br />

connections 299<br />

Eurotium repens 230, 308, 309<br />

euseptate 473<br />

evolution of fungi 35, 133, 165, 221,<br />

246, 511; co-evolution with host<br />

plants 116, 393<br />

exhabitant (of lichens) 446<br />

Exidia glandulosa 603–604; basidium<br />

489, 604<br />

Exobasidium 655–657, Pl. 12; basidium<br />

656, 657; haus<strong>to</strong>rium 656, 657<br />

Exobasidium japonicum 656<br />

Exobasidium vaccinii 656, Pl. 12<br />

Exobasidium vexans 656<br />

Exobasidiales 655–657; biotrophy 655,<br />

656; exocy<strong>to</strong>sis; see secretion<br />

Exophiala 486<br />

extrahaus<strong>to</strong>rial matrix 103, 398, 617<br />

extrahaus<strong>to</strong>rial membrane 617<br />

eyespot (of zoospore) 72<br />

fairy rings 13, 532, 546, 547, 552<br />

famoxadone 113<br />

fatty acids, in biological control 412<br />

fenpropimorph 411<br />

Fen<strong>to</strong>n reaction 530<br />

fibrosin bodies 404, 405<br />

Fibulomyces 689, 690<br />

Filobasidiella neoformans 661; see<br />

Cryp<strong>to</strong>coccus neoformans<br />

fimbriae 176, 202, 275, 277, 653, 654;<br />

fimbrial RNA 654<br />

fingerprint vacuoles, see dense-body<br />

vesicles 76, 107<br />

fish farming 665<br />

fish pathogens, Saprolegnia 81<br />

fission yeast 254<br />

fissitunicate; see ascus wall<br />

flagellar apparatus 129–130, 153<br />

Flagellospora 342<br />

Flagellospora curvula 689, 692<br />

Flagellospora penicillioides (teleom.<br />

Nectria) 689, 692<br />

flagellum 23, 69; retraction 155;<br />

straminipilous (tinsel) flagellum 24,<br />

67–70; whiplash flagellum 68, 69,<br />

153<br />

Flammulina 546<br />

Flammulina velutipes 169, 548–549,<br />

Pl. 9; cultivation 525, 548;<br />

gravitropism 548, 549; oidium 502,<br />

503, 548; trama 524<br />

flexuous hypha 622<br />

flocculation 275<br />

floral mimicry 614<br />

fluconazole 279, 664<br />

5’-fluorocy<strong>to</strong>sine 279, 280, 663<br />

fly agaric 541; see Amanita muscaria<br />

foliose lichens 447, Pl. 8<br />

Fomes fomentarius 560<br />

Fomi<strong>to</strong>psis pinicola 529<br />

food production 167, 181, 302–303;<br />

see cultivation of mushrooms<br />

food spoilage 298; canned food 307<br />

foot cell, in Aspergillus 308, 309; in<br />

Blumeria 394<br />

form-genus 23<br />

formae speciales 120, 345, 346, 395, 621<br />

fosetyl-Al 113, 119<br />

fossil fungi 35, 133, 203, 221, 246, 363,<br />

512, 526, 579; lichens 454<br />

free gills 523<br />

French bean rust 631<br />

fruity smells 371, 373<br />

fruticose lichens 447, Pl. 8<br />

fucosterol 90<br />

Fuligo septica 51, Pl. 1<br />

fumonisin 348<br />

fungi, key features 1–2<br />

fungicide resistance 279–280, 410,<br />

439; see ABC transporters<br />

fungicides, against Phy<strong>to</strong>phthora 112,<br />

113; against powdery mildews 410,<br />

411–412; see also allyl amines;<br />

2-aminopyrimidine, azoles,<br />

benzimidazole, Bordeaux mixture,<br />

captan, carboxin, copper,<br />

dithiocarbamate, imidazoles,<br />

maneb, myclobutanil, morpholine,<br />

phenylamide, phosphonate,<br />

quinoxyfen, strobilurin,<br />

sulphur dust and lime, triazole<br />

fungistasis 534<br />

funiculus 581, 584<br />

furcate basidium 598, 599–600<br />

Furia 203<br />

Furia americana 217, 219, Pl. 3<br />

Fusarium 30, 342–343, 344, 345–348;<br />

biological control 347; endophytes<br />

347; macroconidium 343, 344, 347;<br />

microconidium 343, 344;<br />

myco<strong>to</strong>xins 348; plant diseases 346;<br />

wilts 345, 346<br />

Fusarium culmorum 347<br />

Fusarium heterosporum Pl. 5<br />

Fusarium moniliforme (teleom.<br />

Gibberella fujikuroi) 339<br />

Fusarium oxysporum 345–347<br />

Fusarium venenatum 339<br />

fusion biotroph 191<br />

fusion septum 176<br />

Gabarnaudia (teleom. Sphaeronaemella)<br />

372<br />

Gaeumannomyces graminis 385, 386, 595<br />

Galac<strong>to</strong>myces 281–282<br />

Galac<strong>to</strong>myces candidus 4, 13, 30, 281<br />

Galac<strong>to</strong>myces geotrichum; see G. candidus<br />

gamma-particles 154<br />

gametangio-gametangiogamy 132,<br />

133, 229<br />

gametangio-game<strong>to</strong>gamy 132, 229<br />

gametangium 177; hypogynous versus<br />

epigynous 156, 157, 159, 163<br />

game<strong>to</strong>gamy: isogamy, anisogamy,<br />

oogamy 132<br />

Gamsylella 675, 676; induction of<br />

nema<strong>to</strong>de traps 683<br />

Ganoderma 564<br />

Ganoderma adspersum 565<br />

Ganoderma applanatum 560, 561,<br />

564, 565<br />

Ganoderma lucidum 561, 564<br />

gasterocarp 577<br />

gasteromycetation 578<br />

gasteromycetes 575, 577–592, Pl. 11;<br />

bole<strong>to</strong>id clade 586–589; dispersal<br />

580; euagarics clade 580, 581–585;<br />

evolution 578–579; gomphoidphalloid<br />

clade 589–592; life cycle<br />

577; phylogeny 579, 580;<br />

propagules 580<br />

Gastroboletus 578<br />

Gastrosuillus 578<br />

Gastrosuillus laricinus 578<br />

Geastrum, gomphoid-phalloid clade<br />

579<br />

Geastrum triplex 588, 589<br />

gemma 30, 590


826 INDEX<br />

gene conversion 319, 423<br />

gene-for-gene hypothesis 114, 118, 383,<br />

398, 408, 477, 484, 619, 642, 649<br />

generative hypha 518<br />

genet 558, 586<br />

Geniculosporium 333, 337, 424<br />

Genistellospora homothallica 225<br />

Geoglossum 443<br />

Geolegnia 93<br />

Geosiphon pyriforme 221<br />

Geotrichum candidum (teleom.<br />

Galac<strong>to</strong>myces candidus) 235<br />

germination, au<strong>to</strong>-inhibi<strong>to</strong>rs 378,<br />

387, 614<br />

germination patterns, variability<br />

208, 210, 214, 496; in<br />

Heterobasidiomycetes 600, 602, 608<br />

germ pore (of spores) 26, 27, 239,<br />

614, 615<br />

germ slit (of spores) 239, 332–333, 479<br />

germ sporangium 178<br />

Gibberella-Fusarium complex 345<br />

Gibberella fujikuroi 339<br />

Gibberella zeae, turgor pressure in<br />

ascus 242<br />

gibberellins 339<br />

Gigaspora decipiens 218<br />

Gilbertella persicaria 172, 177, 179<br />

gills; see lamellae<br />

girdling 342, 376, 634<br />

gleba 424, 425, 577, 581, 588, 591<br />

Gliocladium 340, 341<br />

glio<strong>to</strong>xin 339<br />

gloeocystidium 566, 572<br />

gloeoplerous hypha 518, 568<br />

Gloeotulasnella cystidiophora 595<br />

Glomales 29, 37, 217–222<br />

Glomerella 387–388<br />

Glomerella cingulata (anam.<br />

Colle<strong>to</strong>trichum gloeosporioides) 387<br />

Glomerellaceae 316, 386, 389<br />

Glomeromycota 221<br />

Glomus mosseae 218, 220, 221<br />

glucans, in cell walls 5, 6,75<br />

glucan matrix 15, 438<br />

glycerol, in turgor generation 381<br />

glycogen (s<strong>to</strong>rage reserve) 28, 76,<br />

381, 392<br />

Golgi apparatus 11, 66–67, 71, 75,<br />

165, 272<br />

Golovinomyces 403<br />

Golovinomyces cichoracearum 403;<br />

resistance breeding 408<br />

Golovinomyces cucurbitacearum 404<br />

Golovinomyces orontii 403<br />

Gomphidius 558<br />

gomphoid-phalloid clade 575–576;<br />

phylogeny 516–517; gasteromycetes<br />

579, 588–592<br />

Gonapodya 127, 162<br />

Gonapodya prolifera 162<br />

gooseberry, powdery mildew 390, 404<br />

grape-vine, bunch rot (grey mould)<br />

436; dead-arm 375; downy mildew<br />

119; powdery mildew 390<br />

Graphium 366, 368, 372<br />

Graphium penicillioides 372<br />

Graphium ulmi 366, 367<br />

grasses, choke disease 355; crown rust<br />

628; endophytes 357–360; powdery<br />

mildew 393–401<br />

gravitropism 169, 522, 548, 549<br />

grazing animals, fescue <strong>to</strong>xicosis 358;<br />

lupinosis 375; mycoses 295;<br />

ryegrass staggers 359<br />

‘green bridge’ 401, 408, 627<br />

green fluorescent protein (GFP) 263<br />

greenhouse gases 532<br />

grevillins 557, 558<br />

grey mould 434, 436<br />

griseofulvin 293, 304<br />

gut-stage cells of Basidiobolus 206<br />

Gymnoascus reessii 295, 296<br />

gymnocarpic hymenophore 520, 521<br />

gymnohymenial; see apothecium<br />

Gymnopus 546<br />

Gymnosporangium 631–632<br />

Gymnosporangium clavariiforme 632<br />

Gymnosporangium cornutum 632<br />

Gymnosporangium fuscum 613, 629,<br />

631–632, Pl. 12<br />

Gymnosporangium juniperi-virginianae<br />

632<br />

gymnothecium 245, 246, 285, 289,<br />

293, 294, 297<br />

Gyrodon 579<br />

Gyromitra esculenta 415<br />

gyromitrin 415<br />

H þ ATPases 12, 14, 221, 398, 617,<br />

618<br />

hallucinogenic mushrooms 541, 546;<br />

see ergotism<br />

halo; see host defence mechanisms<br />

hamathecium 241<br />

hapteron 584<br />

Hap<strong>to</strong>glossa 54, 63, 64–66, 681; gun cell<br />

64, 65, 66; taxonomy 54<br />

hap<strong>to</strong>r 205<br />

Harpella melusinae 223, 224, 225;<br />

asexual reproduction 223–225;<br />

sexual reproduction 225<br />

Harpellales 223; phylogeny 166, 225<br />

Harpochytrium 162<br />

Harposporium anguillulae 349, 681, 682<br />

Hartig net 21, 526, 527, 586<br />

hart’s truffle 314<br />

haus<strong>to</strong>rial branches 605, 606<br />

haus<strong>to</strong>rial mother cell 615<br />

haus<strong>to</strong>rium 103, 105, 118, 381, 399,<br />

615–616, 617–618; intraparietal<br />

452; see Albugo candida, Blumeria<br />

graminis, lichens, Peronospora,<br />

Phy<strong>to</strong>phthora infestans, Pip<strong>to</strong>cephalis,<br />

Uredinales<br />

hazel, powdery mildew 406<br />

HC-<strong>to</strong>xin 476, 477<br />

heavy metal accumulation; mercury<br />

536<br />

Helicodendron 697; ecophysiology<br />

699–701<br />

Helicodendron conglomeratum 698, 701<br />

Helicodendron fractum 701<br />

Helicodendron fuscosporum 701<br />

Helicodendron giganteum 699<br />

Helicodendron hyalinum 701<br />

Helicodendron praetermissum 701<br />

Helicodendron triglitziense 698, 701<br />

Helicodendron tubulosum 701<br />

Helicoon 697<br />

Helicoon richonis 697, 698<br />

helicospore 23<br />

Heliscus 342<br />

Heliscus lugdunensis (teleom. Nectria<br />

lugdunensis) 685, 687<br />

Helminthosporium 30–31, 471<br />

Helminthosporium velutinum 472, 473<br />

Helotiales (inoperculate<br />

discomycetes) 429–445, Pl. 7;<br />

classification 429, 430; lichens<br />

455<br />

Helvella 423; ec<strong>to</strong>mycorrhiza 423<br />

Helvella crispa 423, Pl. 6<br />

Helvella lacunosa 423<br />

hemiangiocarpic 545; see apothecium,<br />

basidiocarp development<br />

Hemiascomycetes 261–284; versus<br />

Archiascomycetes 261–262;<br />

importance 262; isolation 262;<br />

occurrence 262<br />

hemibiotrophy 388, 475, 550<br />

hemicellulase 530


INDEX<br />

827<br />

hemicellulose 528<br />

Hemileia vastatrix 632–634; longdistance<br />

spread 633<br />

Hemitrichia serpula 51<br />

het genes 325–326, 330<br />

Heterobasidiomycetes 593–608, Pl. 11;<br />

parenthesome 499, 593; taxonomy<br />

593, 594<br />

heterobasidiomycete yeasts 659,<br />

660–666<br />

Heterobasidion abietinum 568<br />

Heterobasidion annosum 519, 567;<br />

blas<strong>to</strong>conidium (Spiniger) 504, 568;<br />

control 569<br />

Heterobasidion parviporum 568<br />

heterobasidium 593;<br />

see phragmobasidium<br />

heteroecious 613<br />

heterogenic incompatibility;<br />

see vegetative incompatibility<br />

heterokaryosis 351, 359; restricted<br />

versus unrestricted 330<br />

heterokont zoospore 24, 71<br />

heterothallism 174, 330, 506–507;<br />

bipolar 506; modified tetrapolar<br />

605; physiological 435; tetrapolar<br />

507<br />

1,16-hexadecanediol 384<br />

hilar appendix (of basidiospore) 27,<br />

490, 491, 492<br />

hilar appendix body 491–492<br />

hilum (of basidiospore) 27, 490<br />

Hirsutella 360<br />

Hirsutella rhossiliensis, as nema<strong>to</strong>de<br />

endoparasite 685<br />

Hirsutella sinensis (teleom. Cordyceps<br />

sinensis) 362<br />

His<strong>to</strong>plasma capsulatum 289, 290–292,<br />

307<br />

Hohenbuehelia 542; nema<strong>to</strong>de-trapping<br />

679, 681<br />

holdfast 223, 224<br />

Holliday model 319, 643<br />

holobasidium 488<br />

holoblastic conidiogenesis 31, 232<br />

holocoenocytic 569<br />

holomorph 32<br />

holothallic conidiogenesis 235<br />

homing reaction (in basidiomycete<br />

fertilization) 502, 503<br />

homing sequence (in zoospores) 100<br />

Homobasidiomycetes 514–564,<br />

565–576, Pl. 9, Pl. 10, Pl. 11; eightclade<br />

system 514, 517;<br />

hymenophore arrangement 514,<br />

515; phylogeny 516<br />

homogenic incompatibility 325<br />

homothallism 319–320, 330, 506;<br />

primary 506; secondary<br />

(¼ pseudohomothallism) 179,<br />

506–507, 535; unclassified 506<br />

honey bees, chalk brood disease 288<br />

honeydew 349<br />

hook cell; see crozier<br />

hops, powdery mildew 404<br />

horizontal gene transfer 305<br />

horizontal resistance; see host<br />

resistance<br />

horizontal transmission of inoculum<br />

357<br />

Hormonema dematioides 227<br />

hormones; see antheridiol, cAMP,<br />

conidiogenone, oogoniol, parisin,<br />

sirenin, trisporic acid; see also plant<br />

growth hormones, sex hormones<br />

host defence mechanisms 397; halo<br />

397; papilla 397, 408;<br />

see hypersensitive response,<br />

oxidative burst, PR proteins<br />

host differentials 625, 628, 642<br />

host-jumping 362<br />

host-pathogen recognition;<br />

see gene-for-gene hypothesis<br />

host resistance; adult plant resistance<br />

410, 620; field (horizontal, partial,<br />

minor gene) resistance 115, 408,<br />

619; major gene (vertical) resistance<br />

114, 408, 619, 626<br />

hot water treatment (of seeds) 651<br />

Hülle cells 308, 310<br />

human pathogens 196, 211, 471;<br />

see Aspergillus flavus, Aspergillus<br />

fumigatus, Blas<strong>to</strong>myces dermatitidis,<br />

Candida albicans, Coccidioides immitis,<br />

Cryp<strong>to</strong>coccus neoformans,<br />

His<strong>to</strong>plasma capsulatum,<br />

Paracoccidioides brasiliensis,<br />

Penicillium marneffei, Pneumocystis,<br />

Trichosporon<br />

Humicola insolens; see Scytalidium<br />

thermophilum<br />

hybrid ascus 319<br />

hybrid vigour 477<br />

Hydnangium 578<br />

Hydnotrya 423<br />

Hydnum repandum 515, 575<br />

Hydnum rufescens 574, 575<br />

hydrogenosome 151<br />

hydrophobins 8, 327, 364, 366, 452,<br />

519, 535, 545<br />

Hygrocybe 546<br />

Hygrocybe coccinea 546, Pl. 9<br />

Hygrocybe conica 546<br />

Hygrocybe psittacina 546<br />

Hygrophoropsis aurantiaca 501<br />

Hygrophorus 546<br />

hymenium 525<br />

Hymenoascomycetes 315<br />

hymenochae<strong>to</strong>id clade 573–574;<br />

phylogeny 516–517<br />

Hymenomycetes 513<br />

hymenophoral trama 523, 524<br />

hymenophore 515<br />

Hymenoscyphus ericae 444<br />

Hymenoscyphus splendens 688<br />

Hymenoscyphus tetracladius 688<br />

Hymenostilbe 360<br />

hyperplasia 56<br />

hypersensitive response 62, 115, 400,<br />

408, 618; in nectrotrophs 438–439,<br />

476<br />

hypertrophy 56, 58, 60–61, 136, 137,<br />

547, 550, 644<br />

hypha 2, 4; in Ascomycota 227; in<br />

Basidiomycota 501; in Oomycota<br />

75–76; Spitzenkörper 3, 4–5, 7, 10;<br />

vessel hypha 17<br />

hyphal analysis 517, 518, 519, 562;<br />

see momomitic, dimitic, sarcomitic,<br />

sarcotrimitic, trimitic<br />

hyphal body 203, 216, 364<br />

hyphal branching 13<br />

hyphal coiling 340, 342<br />

hyphal interference 540, 569<br />

hyphal tip 3, 4–5, 8; and Ca 2þ gradient<br />

9–10<br />

Hyphochytriomycota 24, 70, 71<br />

Hyphochytrium catenoides 70<br />

Hyphochytrium peniliae 71<br />

Hypholoma fasciculare 554<br />

Hypholoma sublateritium 553, 554<br />

hyphomycetes 231<br />

hyphopodium 386<br />

hyphosphere 534<br />

Hypocrea 339–341<br />

Hypocrea gelatinosa 340, 341<br />

Hypocrea jecorina (anam. Trichoderma<br />

reesei) 340<br />

Hypocrea poronioidea 341<br />

Hypocrea pulvinata 339, 340, Pl. 5<br />

Hypocrea rufa (anam. Trichoderma viride)<br />

340


828 INDEX<br />

Hypocrea virens (anam. Trichoderma<br />

virens) 340<br />

Hypocreales 316, 337–348; anamorphs<br />

316, 338, 339; lichens 455;<br />

taxonomy 339<br />

hypogeous fruit bodies 313, 414,<br />

423<br />

hypothallus, of Myxomycetes 49, 52<br />

hypovirulence 368, 376–377<br />

Hypoxylon 335–336, 607<br />

Hypoxylon fragiforme 333, 335–336,<br />

337; eclosion of ascospores 338<br />

Hypoxylon mammatum 333<br />

Hypoxylon multiforme 337<br />

Hypoxylon rubiginosum 335<br />

Hypovirus 376–377; CHV-1 genome<br />

376; transmission 377<br />

hysterothecium 459<br />

ibotenic acid 539, 541<br />

idiomorphs 267; see mating-type<br />

imidazole-type fungicides 479<br />

imperfect state; see anamorph<br />

inaequi-hymenial gills 522<br />

incertae sedis 289<br />

incompatibility, homogenic versus<br />

heterogenic 325<br />

indole 630<br />

infection bulb 684<br />

infection hypha; bulbous or<br />

primary 378, 381, 388; secondary<br />

388<br />

infection mechanisms; see Blumeria<br />

graminis, Hap<strong>to</strong>glossa, Magnaporthe<br />

grisea, nema<strong>to</strong>phagous fungi,<br />

Plasmodiophora brassicae, Uredinales<br />

infection plaque 440, 441<br />

Ingoldiella hamata (teleom.<br />

Sis<strong>to</strong>trema hamatum) 243, 689, 691<br />

inhabitant (of lichens) 446<br />

ink-caps 537; see Coprinus<br />

Inonotus 573<br />

inoperculate discomycetes 429;<br />

see Helotiales, Lecanorales<br />

insect associations; see Amylostereum,<br />

Epichloe, Ganoderma, Laboulbeniales,<br />

Trichomycetes<br />

insect dispersal and fertilization 290,<br />

293, 538, 571, 580, 654; in Claviceps<br />

349; in Epichloe 356; in Ophios<strong>to</strong>ma<br />

365; Phallus 590–591; in Uredinales<br />

621, 630<br />

insect diseases; muscardine disease<br />

362–363<br />

insect pathogens; see Ascosphaera apis,<br />

Beauveria bassiana, Conidiobolus<br />

coronatus, Cordyceps, En<strong>to</strong>mophthora<br />

muscae, Erynia neoaphidis, Harpella<br />

melusinae, Metarhizium anisopliae; of<br />

mosqui<strong>to</strong> and black fly larva<br />

see Coelomomyces, Erynia conica,<br />

Lagenidium giganteum<br />

integrin 9, 615<br />

inter-costal veins 27, 327<br />

interference competition; see hyphal<br />

interference<br />

internal proliferation 83, 99, 144<br />

invasive growth 8<br />

investing hyphae 416<br />

involutin 555, 557<br />

g-irradiation resistance 417<br />

isidium 448<br />

isozyme analysis 34<br />

isthmus 300<br />

Itersonilia perplexans 493; ballis<strong>to</strong>spore<br />

discharge 494–495<br />

ITS (internal transcribed spacer) 35<br />

itraconazole 306–307<br />

jelly fungi 593, Pl. 11<br />

keratinolytic fungi 102, 293<br />

Kickxellales 166, 225<br />

killer <strong>to</strong>xins 273–274, 646–647<br />

killer yeasts 273<br />

kine<strong>to</strong>some 68, 130, 162; dormant<br />

kine<strong>to</strong>some 71, 129, 162<br />

kresoxim methyl 410, 412<br />

Kretzschmaria deusta 333<br />

Laboulbeniales 226<br />

Labyrinthula 72, 73<br />

Labyrinthulomycota 71–72, 73, 74<br />

Laccaria 552; ec<strong>to</strong>mycorrhiza 552;<br />

related <strong>to</strong> gasteromycetes 578<br />

Laccaria amethystina 552<br />

Laccaria bicolor 552<br />

Laccaria laccata 552<br />

laccase 439, 530<br />

Lactarius 566<br />

Lactarius deliciosus 566, Pl. 10<br />

Lactarius deterrimus 566<br />

Lactarius lignyotellus 491<br />

Lactarius pyrogalus 566<br />

Lactarius quietus 566<br />

Lactarius rufus 519, 566<br />

Lactarius <strong>to</strong>rminosus 566<br />

Lactarius turpis 566<br />

lactifer (lacticiferous hypha) 519, 566,<br />

568, 569<br />

Laetiporus 564<br />

Laetiporus sulphureus 504, 505, 519,<br />

562, 564, Pl. 10<br />

Lagenidium giganteum 79<br />

Lalaria 252, 658<br />

lamellae 522–523; attachment <strong>to</strong> stipe<br />

523; see aequi-hymenial, inaequihymenial<br />

Langermannia; see Calvatia<br />

latent infections 375, 388, 628<br />

latex 551<br />

Lecanora 455<br />

Lecanora conizaeoides 450, 454–455<br />

Lecanora (Sphaerothallia) esculenta 455<br />

Lecanora muralis 455, Pl. 8<br />

Lecanorales 446–458; rostrate ascus<br />

dehiscence 455; see lichens<br />

lecanoric acid 453<br />

Leccinum 558<br />

Leccinum scabrum 558<br />

Leccinum versipelle 558<br />

lectins 175, 683<br />

legislation (in plant diseases) 138, 649<br />

legumes, Ascochyta diseases 464; leafspot;<br />

see also individual legumes 466<br />

Lemonniera aquatica 685, 687<br />

Lemonniera terrestris 685<br />

Lentinula 546<br />

Lentinula edodes; cultivation 525<br />

Lentinus tigrinus 521; secotioid form 578<br />

Leotia lubrica 444, Pl. 7<br />

Leotiales; see Helotiales<br />

Lepista 552<br />

Lepista nuda 552<br />

Lepista saeva 552<br />

leprose lichens 447<br />

Lep<strong>to</strong>graphium 366<br />

Lep<strong>to</strong>sphaeria 460–464, 472; anamorphs<br />

461; versus Phaeosphaeria 462–464<br />

Lep<strong>to</strong>sphaeria acuta (anam. Phoma acuta)<br />

231, 461, 462, 463; endophytic stage<br />

462<br />

Lep<strong>to</strong>sphaeria biglobosa 463<br />

Lep<strong>to</strong>sphaeria coniothyrium 461<br />

Lep<strong>to</strong>sphaeria maculans (anam. Phoma<br />

lingam) 461, 463–464<br />

lethal reactions, in Physarum<br />

polycephalum 48, 52<br />

Letharia vulpina 452<br />

lettuce, big-vein virus (LBVV) 148;<br />

downy mildew 120<br />

Leveillula 407


INDEX<br />

829<br />

Leveillula taurica 407<br />

Lewia (anam. Alternaria) 460, 469<br />

Lewia infec<strong>to</strong>ria 469, 470<br />

Liceales (Myxomycota) 51<br />

lichen acids 453<br />

lichen desert 454<br />

lichenicolous fungi 451<br />

lichenicolous lichens 451<br />

lichenometry 447<br />

lichens 446–458, Pl. 8; air pollution<br />

454; chimaera 457; haus<strong>to</strong>rium 450,<br />

452; nutrition 451–452; phylogeny<br />

226, 454; pigments 452, 453, 454;<br />

pure-culture studies 446;<br />

reproduction 448; secondary<br />

metabolism 453; stratification 447,<br />

448; taxonomy 448, 449, 454–455;<br />

thallus establishment 448–449, 450,<br />

451; thallus morphology (crus<strong>to</strong>se,<br />

foliose, fruticose) 447–448<br />

lignin 528; degradation 530–532<br />

lignin peroxidase 530, 531<br />

Limnoperdon 577<br />

linoleic acid, nematicidal properties<br />

683<br />

lipid (s<strong>to</strong>rage reserves) 28–29, 71, 78,<br />

86, 350, 381, Pl. 12; in oospore<br />

arrangement (Oomycota) 86<br />

lipid sac 154<br />

locule 315, 459<br />

Loculoascomycetes 315, 459–486<br />

loline 359, 360<br />

lolitrem 359, 360<br />

long-distance dispersal 400, 586, 610,<br />

624, 633, 639<br />

Lophodermiella 441<br />

Lophodermium 441<br />

LSD 554<br />

Lunulospora curvula 691, 692<br />

lupinosis 375<br />

Lycogala epidendron 51, Pl. 1<br />

Lycoperdon 581<br />

Lycoperdon perlatum 582<br />

Lycoperdon pyriforme 581, 583<br />

Lyophyllum 552<br />

lysergic acid 354<br />

lysigenous perithecial development<br />

315, 459<br />

lysine biosynthesis, DAP and AAA<br />

pathways 67, 71, 75, 165<br />

macroconidia; see Botrytis cinerea,<br />

Fusarium, Neurospora, Sclerotinia<br />

macrocyclic (rusts) 613<br />

macrocysts 42<br />

Macrolepiota 536; related <strong>to</strong><br />

gasteromycetes 579, 581<br />

Macrolepiota procera 533, 536; and<br />

mercury accumulation 536<br />

Macrolepiota rhacodes 536<br />

macronema<strong>to</strong>us 230<br />

macrophages 291<br />

macrophore (in Phycomyces) 168<br />

Magnaporthaceae 316, 377–386<br />

Magnaporthe grisea 378, 379–382, 383,<br />

384–385, 388, 396; appressorium<br />

formation 378–381; appressorium<br />

maturation 381; conidium<br />

germination 378; gene-for-gene<br />

relationship 383; penetration 381;<br />

signalling 384, 385<br />

Magnusiomyces 282<br />

maize, anthracnose 388; cms-T<br />

mutation 477; grey leaf spot 481;<br />

Northern leaf spot 477; root and<br />

stalk rot 95; rust 630; smut 643;<br />

Southern leaf blight 477<br />

major gene resistance; see host<br />

resistance<br />

Malassezia 671–672; dandruff 672;<br />

ultrastructure 671, 672<br />

Malassezia furfur 671<br />

Malassezia globosa 672<br />

Malassezia pachydermatis 671, 672<br />

Malasseziales 671–672<br />

Malbranchea 289, 290, 292, 295<br />

maneb 113, 631, 651<br />

manganese peroxidase 530<br />

manna lichen 455<br />

mantle (of ec<strong>to</strong>mycorrhiza) 21, 526, 527<br />

MAP kinase 384, 385, 645<br />

Marasmius 18, 546–548<br />

Marasmius androsaceus 548<br />

Marasmius oreades 511, 546, 547<br />

Marasmius ramealis 548<br />

Marasmius rotula 548<br />

Margaritispora aquatica 685, 691, 693<br />

Massarina 688, 690<br />

Massarina aquatica 693<br />

Massospora 203<br />

mastigoneme; see tripartite tubular<br />

hair<br />

mating behaviour, Ascomycota 229,<br />

266–270, 278, 322–324, 330–331;<br />

Basidiomycota 506–510;<br />

Chytridiomycota 131–133;<br />

Oomycota 101–102; Zygomycota<br />

173–181<br />

mating-type, Ascomycota: idiomorphs<br />

277, 320, 330, 436, 464, 469, 477;<br />

loci 268; multiple alleles 387;<br />

switch 268, 269, 270, 277<br />

mating-type, Basidiomycota: A and B<br />

loci 508, 509–510, 662; heterodimer<br />

formation 509, 645; multiple alleles<br />

507; structure 508–510<br />

medicinal fungi 361, 562, 565, 608<br />

Megacollybia platyphylla 500, 502, 511<br />

meiosporangium, meiospore 156, 176<br />

Melampsora 635, Pl. 12<br />

Melampsora epitea 635<br />

Melampsora euphorbiae, telium 635<br />

Melampsora lini var. lini 635;<br />

gene-for-gene concept 619<br />

Melampsoridium betulinum 635<br />

melanin 16, 177, 298, 381, 664, Pl. 4;<br />

DHN melanin biosynthesis 381;<br />

DOPA melanin 664<br />

Melzer’s iodine 28, 241–242, 566<br />

membrane channels (stretchactivated);<br />

see Ca 2þ<br />

membrane cycling 12, 272<br />

memnospore 22, 29<br />

meningoencephalitis 663<br />

Meripilus giganteus 501<br />

meristem arthroconidium 390, 391<br />

meris<strong>to</strong>spore 205, 206<br />

merosporangium 166, 195, 201<br />

mesospore 629<br />

metabasidium 490, 611, 637<br />

metalaxyl 113, 119, 126<br />

Metarhizium anisopliae 361, 363, 364;<br />

penetration 364<br />

methylhydrazine 415<br />

metula 298, 310, 311<br />

Microascales 316, 368–373<br />

Microascus 368–369; perithecium<br />

formation 369<br />

microbodies 71, 684<br />

microbody-lipid complex (MLC) 130,<br />

131<br />

Microbotryales 609, 636, 652–655;<br />

intercellular growth 652; life cycle<br />

652; phylogeny 652<br />

Microbotryum 639; dicot Ustilago 652<br />

Microbotryum violaceum 652–655;<br />

conjugation 653; ecology 655;<br />

fimbriae 653, 654; host-pathogen<br />

interactions 654–655; life cycle 652;<br />

sex chromosomes 638, 653; somatic<br />

diploids; systemic infection 655;<br />

teliospore germination 653


830 INDEX<br />

microconidium 291, 329, 343,<br />

346, 435<br />

microcyclic (rusts) 614<br />

microcyst 41–42<br />

microfilaments; see actin<br />

Micromucor 197<br />

micronema<strong>to</strong>us 230<br />

microphore (in Phycomyces) 168<br />

Microsphaera 401–403<br />

Microsphaera (Erysiphe) alphi<strong>to</strong>ides 390,<br />

402, 403<br />

Microsporum 235, 293, 294<br />

microtubules 9–10, 79, 106, 158, 258,<br />

270, 615, 646; arrangement in<br />

flagella 68; ascospore formation<br />

323; nuclear migration 499<br />

millet, ergot 349<br />

Minimedusa polyspora 505<br />

mint rust 613<br />

mirror yeast 666<br />

mites, dispersal of fungi 448<br />

mi<strong>to</strong>chondria 4, 155, 264; cristae of<br />

inner membrane 67, 68, 71, 165<br />

mi<strong>to</strong>gen 385<br />

mi<strong>to</strong>sporangium, mi<strong>to</strong>spore 159<br />

mi<strong>to</strong>sporic fungi 32<br />

mlo allele (barley) 397, 408, 475<br />

Moesziomyces penicillariae 652<br />

molecular clock 35<br />

Mollisia 439<br />

Mollisia cinerea 439, Pl. 7<br />

Mollisia uda 691<br />

Monilia 432<br />

Monilinia fructicola 432<br />

Monilinia fructigena 31, 430, 432, 433;<br />

sclerotium 432<br />

Monilinia laxa 432<br />

Monilinia polystroma 432<br />

Moniliopsis (teleom. Thanatephorus) 594<br />

Monoblepharidales 162–164; zoospore<br />

162, 163<br />

Monoblepharella 162<br />

Monoblepharis 162–164; sexual<br />

reproduction 163<br />

Monoblepharis macrandra 163, 164<br />

Monoblepharis polymorpha 163<br />

Monoblepharis sphaerica 164<br />

monoclinous antheridium 85<br />

monomitic basidiocarp construction<br />

519<br />

Monosporascus cannonballus 237<br />

monoverticillate; see Penicillium<br />

conidiophores<br />

Montagnea 578<br />

Morchella 427–428; cultivation 428;<br />

ec<strong>to</strong>mycorrhiza 427–428;<br />

endobacteria 428; saprotrophic<br />

populations 427; sclerotia 428<br />

Morchella angusticeps 427<br />

Morchella conica 427<br />

Morchella crassipes 427<br />

Morchella deliciosa 427<br />

Morchella elata 427<br />

Morchella esculenta 427, Pl. 6<br />

Morchella semilibera 427<br />

morpholine-type fungicides 410<br />

morphological heterothallism 200<br />

Mortierella 197–200; zygospores 198,<br />

200<br />

Mortierella alpina 197<br />

Mortierella capitata 200<br />

Mortierella chlamydospora 197<br />

Mortierella erice<strong>to</strong>rum 200<br />

Mortierella humilis 197<br />

Mortierella hyalina 198<br />

Mortierella indohi 199<br />

Mortierella isabellina 197<br />

Mortierella polycephala 197<br />

Mortierella ramanniana 197<br />

Mortierella rostafinskii 199<br />

Mortierella stylospora 197<br />

Mortierella umbellata 198, 200<br />

Mortierella (Umbelopsis) vinacea 25,<br />

197<br />

Mortierella wolfii 197<br />

Mortierella zonata 197, 199<br />

mucilage, in yeasts 659, 661; (in<br />

attachment) see adhesives<br />

Mucor 180–181<br />

Mucor azygospora 179, 181<br />

Mucor bainieri 179<br />

Mucor circinelloides 181<br />

Mucor fragilis 181<br />

Mucor genevensis 175, 179<br />

Mucor hiemalis 172–173, 176, 179, 181;<br />

azygospore 179<br />

Mucor indicus 181<br />

Mucor lusitanicus 181<br />

Mucor miehei 181<br />

Mucor mucedo 167, 172, 179<br />

Mucor piriformis 173, 179<br />

Mucor plasmaticus 173<br />

Mucor plumbeus 30, 173, 182, 189<br />

Mucor pusillus 178, 181<br />

Mucor racemosus 181, 182<br />

Mucor rouxii, yeast stage 167, 168,<br />

181<br />

Mucor spinosus 181<br />

Mucorales 165–200; homothallism<br />

173; mating behaviour 178–179; sex<br />

hormones 174; sexual reproduction<br />

173–178; sporangium development<br />

171–173<br />

mucormycosis 181, 183<br />

muscimol 539, 541<br />

mushroom poisoning 415, 540,<br />

555–556; Emperor Claudius 540<br />

Mutinus caninus 589, 591–592<br />

mycelial cords 15, 500, 502, 559; in<br />

mycorrhiza 526, 585<br />

Myceliophthora roreri; see Crinipellis<br />

roreri<br />

mycelium 2; dikaryotic 3, 492, 497,<br />

501; heterokaryotic 3, 227, 497;<br />

homokaryotic 2, 227, 496;<br />

monokaryotic 3, 492, 496; mycelial<br />

strands 15; rhizomorphs 16, 17–18<br />

Mycena 166, 551–552; latex 551; orchid<br />

mycorrhiza 598<br />

Mycena galericulata 551<br />

Mycena galopus 551, 570<br />

Mycena pura Pl. 3<br />

Mycena sanguinolenta 551<br />

Mycena stylobates 551<br />

myclobutanil 479<br />

mycobiont (of lichens) 446<br />

Mycocentrospora filiformis 694<br />

mycoheterotrophy 526<br />

mycolaminarin 76, 83, 87, 107<br />

mycoparasitism 97, 127, 191, 546,<br />

556, 605<br />

mycophagous insects 565, 591<br />

mycorrhiza 218; see ec<strong>to</strong>mycorrhiza,<br />

ericoid mycorrhiza, orchid<br />

mycorrhiza, VAM<br />

mycoses, Basidiobolus 207; Fusarium<br />

343; Mucorales 181, 183; Pythium<br />

insidiosum 102; Scopulariopsis 369;<br />

see human pathogens<br />

Mycosphaerella 464, 480–481, 484;<br />

anamorphs 481; diseases 480, 481;<br />

pseudothecium 481, 482<br />

Mycosphaerella dianthicola<br />

(anam. Cladosporium<br />

echinulatum) 482<br />

Mycosphaerella graminicola (anam.<br />

Sep<strong>to</strong>ria tritici) 481, 483; genetic<br />

diversity 481<br />

Mycosphaerella macrospora<br />

(anam. Cladosporium humile) 482<br />

Mycosphaerella tassiana (anam.<br />

Cladosporium herbarum) 482


INDEX<br />

831<br />

mycosporine-alanine 387<br />

myco<strong>to</strong>xins 302, 304, 305–306, 347,<br />

348, 470<br />

Mycotypha microspora 202<br />

Mycovellosiella 483<br />

Mycovellosiella fulva; see Cladosporium<br />

fulvum<br />

mycoviruses 329, 375–376;<br />

transmission 377; see Hypovirus<br />

Myriosclerotinia curreyana 430;<br />

see Myrioconium, Sclerotinia curreyana<br />

myxamoeba 49<br />

Myxomycetes 47–53, Pl. 1<br />

Myxotrichum chartarum 296, 297;<br />

taxonomic position 297, 444<br />

Myzocytium 681<br />

nasse apicale 241<br />

necrotrophy 388, 432, 434, 438<br />

Nectria 341–343, 689<br />

Nectria cinnabarina 341–342, 344, 600,<br />

Pl. 5<br />

Nectria coccinea 341<br />

Nectria galligena (anam. Cylindrocarpon<br />

heteronema) 341, 343, 344, 348<br />

Nectria haema<strong>to</strong>cocca 344<br />

Nectria (anam. Heliscus) lugdunensis 687<br />

Nectria mammoidea 345<br />

Nectria radicicola (anam. Cylindrocarpon<br />

destructans) 348<br />

Nema<strong>to</strong>c<strong>to</strong>nus (teleom. Hohenbuehelia)<br />

679, 681<br />

nema<strong>to</strong>des 673<br />

nema<strong>to</strong>phagous fungi 63, 64, 349,<br />

542, 673–685; adhesion 683;<br />

adhesive knobs 675, 676, 679;<br />

adhesive nets 675, 676; biological<br />

control 684; constricting rings<br />

676–679; egg and cyst parasites 674,<br />

684; endoparasitic fungi 674,<br />

680–682; infection process 683;<br />

non-constricting rings 676;<br />

preda<strong>to</strong>ry fungi 673, 675–680;<br />

taxonomy 674; <strong>to</strong>xins 683; trap<br />

induction 682–683; trophic hyphae<br />

673, 684<br />

Neobulgaria 444<br />

Neocallimastigales 150–153<br />

Neocallimastix 151<br />

Neocallimastix frontalis 151<br />

Neocallimastix hurleyensis 151, 152<br />

Neocallimastix patriciarum 151<br />

Neoerysiphe galeopsidis 404<br />

Neonectria 348<br />

Neotyphodium (teleom. Epichloe) 355,<br />

358; alkaloids 358–359, 360;<br />

mutualistic symbiosis 359; origin<br />

359<br />

Neotyphodium coenophialum 342<br />

Neotyphodium lolii 358, 359<br />

Neotyphodium typhinum 358<br />

Neozygites 203<br />

Neurospora 326–331; as a genetic <strong>to</strong>ol<br />

327; life cycle 327–330<br />

Neurospora africana 327<br />

Neurospora crassa 11, 229, 237, 320,<br />

327, 329; biological clock 331; life<br />

cycle 327, 328, 329–330; mating<br />

behaviour 330–331<br />

Neurospora dodgei 327<br />

Neurospora intermedia 326<br />

Neurospora si<strong>to</strong>phila 327<br />

Neurospora terricola 327<br />

Neurospora tetrasperma, ascospore 26,<br />

228, 237; pseudohomothallism<br />

323–324, 506<br />

neurosporoxanthin 329<br />

Nia vibrissa 577, 579, 693<br />

Nigrospora 30<br />

Nodulisporium 333, 334, 335<br />

Nodulisporium tulasnei 335<br />

Nos<strong>to</strong>c 221, 446, 456<br />

Nowakowskiella 144–145<br />

Nowakowskiella elegans 144, 145<br />

Nowakowskiella profusa 129, 144<br />

nuclei 4, 5, 79, 228, 497, 501, Pl. 12;<br />

conjugate division 499; nuclear<br />

migration 498<br />

numbers of species of fungi 1, 226<br />

numbers of spores 497<br />

nutrient uptake 12–13, 14<br />

nystatin 279<br />

oak, powdery mildew 402, 403; wilt<br />

369<br />

oats, covered smut 643; crown rust<br />

628; leaf necrosis 476; loose smut<br />

642, 643; Vic<strong>to</strong>ria cultivar 476, 619<br />

ocellus 186<br />

ochra<strong>to</strong>xin A 304, 305<br />

odours 371, 373, 424, 591, 630<br />

Oedocephalum 419, 420, 504<br />

Oedogoniomyces 162<br />

Oidiodendron 295, 444<br />

Oidiopsis 408<br />

oidium 421, 502, 503, 538<br />

Oidium 390<br />

Oidium tuckeri 403<br />

oilseed rape, leaf spot 439; blackleg<br />

461, 463<br />

oleaginous fungi 167<br />

Olpidium 145–148; fossil 133;<br />

transmission of viruses 147–148;<br />

zoospores 146<br />

Olpidium bornovanus 146–148<br />

Olpidium brassicae 127, 146–147, 148<br />

Omphalotus 546<br />

Omphalotus olearius 501<br />

one-gene-one-enzyme hypothesis 327<br />

onions, leaf blight 468; white rot 434<br />

Onygena 290, 293<br />

Onygena corvina 290, 293<br />

Onygena equina 290, 293<br />

Onygenales 289–297<br />

oogamous fertilization 28, 77, 162<br />

oogoniol 90<br />

oogonium 29, 77, 86, 89, 101<br />

Oomycota 24, 38, 64, 75–126, P. 2;<br />

hypha 75–76; sexual reproduction<br />

77, 78; taxonomy 80; zoospore 76,<br />

77<br />

ooplast 78, 86, 102, 110<br />

oosphere 29, 102, 109<br />

oospore 28, 29, 77, 102, 109–110,<br />

164; centric, subcentric,<br />

subeccentric, eccentric types 86<br />

oosporein 363<br />

operculate discomycetes 414;<br />

see Pezizales<br />

operculum 241, 414, 417, 418;<br />

see ascus apical apparatus<br />

Ophios<strong>to</strong>ma 364–366; perithecium 364,<br />

365, 366; versus Cera<strong>to</strong>cystis 364, 373<br />

Ophios<strong>to</strong>ma himal-ulmi 366–367<br />

Ophios<strong>to</strong>ma novo-ulmi 366; control of<br />

Dutch elm disease 368<br />

Ophios<strong>to</strong>ma piceae 365<br />

Ophios<strong>to</strong>ma ulmi 366, 367<br />

Ophios<strong>to</strong>matales 316, 364<br />

opisthokont zoospore 24<br />

-opsis form (of Uredinales) 613<br />

Orbilia 675<br />

Orbiliaceae 675<br />

orchid mycorrhiza 550, 596, 597, 598<br />

orellanine 539, 555<br />

organohalogens, fungal origin 530<br />

Orpinomyces 151<br />

Oudemansiella canarii 489<br />

Oudemansiella mucida 548<br />

Oudemansiella radicata 547, 548;<br />

basidium 488<br />

Ovulariopsis 406


832 INDEX<br />

oxalate crystals 173, 529<br />

oxalic acid 433, 438, 447, 529<br />

oxidative burst 291, 397, 438, 475, 540<br />

oxidative stress 434, 664; and<br />

carotenoid production 434, 665,<br />

670, 671<br />

oyster mushroom; see Pleurotus<br />

Ozonium 537<br />

Pachysolen tamophilus 262<br />

Paecilomyces 299, 360<br />

Paecilomyces lilacinus, as nema<strong>to</strong>de egg<br />

parasite 684; biological control 685<br />

Paecilomyces militaris (teleom. Cordyceps<br />

militaris) 361<br />

Palmella stage of Basidiobolus ranarum<br />

206<br />

Panaeolus 546<br />

Panaeolus semi-ovatus 546<br />

Panaeolus sphinctrinus 543, 546<br />

Pandora 203<br />

Panellus stypticus 489, 491<br />

papilla; see host defence mechanisms<br />

Paracoccidioides brasiliensis 289,<br />

290–292<br />

paracrystalline bodies 141<br />

paragynous fertilization 109<br />

paraphysis 241, 315, 414, 461; apical<br />

paraphysis 337, 343, 461; in<br />

Basidiomycota 522, 525<br />

parasexual cycle 115, 230, 345<br />

parasitic fungi; see individual host<br />

plants; algae, insects,<br />

mycoparasitism, mycoses,<br />

nema<strong>to</strong>des, rotifers<br />

Parasola 537<br />

Parauncinula septata 393<br />

parenthesome 497, 498–499, 573, 593,<br />

606<br />

parietin 453, 456<br />

parisin 156<br />

parthenogenesis 85, 179<br />

partial veil 520<br />

pathogenesis-related proteins;<br />

see PR proteins<br />

patulin 304, 306–307<br />

pavement cracking 538<br />

Paxillus 555<br />

Paxillus atro<strong>to</strong>men<strong>to</strong>sus 555<br />

Paxillus involutus 501, 555, 556;<br />

post-meiotic mi<strong>to</strong>sis 490<br />

pea, blight 464; leaf and pod spot 464;<br />

powdery mildew 402<br />

peach, leaf curl 251; brown-rot 432<br />

pea truffle; see Endogone<br />

pear trellis rust 631, Pl. 12<br />

pectin 528<br />

pectinolytic enzymes 433–434,<br />

438, 529<br />

pelo<strong>to</strong>n 597<br />

Peltigera 456<br />

Peltigera canina 456, Pl. 8<br />

Peltigera polydactyla 456, 457<br />

penetration peg 381, 396<br />

penicillin 303, 304<br />

Penicillium 297–299, 310–313;<br />

conidiophore 298, 299; mono-,<br />

bi-, ter-, quaterverticillate 298, 310,<br />

311; sclerotium 313<br />

Penicillium aurantiogriseum 312<br />

Penicillium camemberti 303, 311<br />

Penicillium chrysogenum 303<br />

Penicillium citrinum 311<br />

Penicillium claviforme 230<br />

Penicillium cyclopium 301<br />

Penicillium digitatum 311–312, Pl. 4<br />

Penicillium expansum 306, 311, 312<br />

Penicillium griseofulvum 293, 304, 311<br />

Penicillium italicum 311–312<br />

Penicillium janczewskii 304<br />

Penicillium marneffei 307; dimorphism<br />

301<br />

Penicillium notatum 303<br />

Penicillium roqueforti 303, 311<br />

Penicillium spinulosum 311<br />

Penicillium verrucosum 305<br />

Penicillium verruculosum 311<br />

Peniophora 605<br />

Peniophora gigantea; see Phlebiopsis<br />

gigantea<br />

Peniophora quercina 515<br />

pentachloronitrobenzene 651<br />

peramine 359, 360<br />

percurrent conidiogenesis 469<br />

perfect state; see teleomorph<br />

periclinal thickening 232<br />

Peridermium 634; see Cronartium<br />

Peridermium pini 635<br />

peridiole 577, 583, 584, 586, 589, 590<br />

peridium 314, 424, 577, 587, 622<br />

perifungal membrane 597<br />

Périgord truffle 426<br />

periphysis 243, 315, 621<br />

periplasm (in oosphere formation) 86,<br />

109<br />

periplasmic space 270<br />

perispore, perisporium (of spore<br />

walls) 26, 237, 417, 420<br />

perithecial hairs 331, 332<br />

perithecial stroma 246, 335,<br />

336–337, 348, 352–353, 360, 362,<br />

373, Pl. 5<br />

perithecium 245, 246, 317;<br />

development 315; in Microascus 369;<br />

in Nectria 337; in Ophios<strong>to</strong>ma<br />

364–366; in Sordaria 318–319; in<br />

Xylaria 333<br />

Peronosclerospora 125<br />

Peronosclerospora sorghi 125–126<br />

Peronospora 116–119; control 119;<br />

haus<strong>to</strong>rium 117–118; oospores 118;<br />

sporangiophore 117<br />

Peronospora destruc<strong>to</strong>r 116, 119<br />

Peronospora farinosa 116<br />

Peronospora manshurica 118<br />

Peronospora parasitica 116, 117,<br />

118–119, 122, Pl. 2<br />

Peronospora tabacina 116, 119<br />

Peronospora trifoliorum 119<br />

Peronospora viciae 119<br />

Peronosporales (downy mildews)<br />

115–125; biotrophy 116; life cycle<br />

116; origin 116; relationships with<br />

Pythiales 115<br />

peroxisome 684<br />

Pertusaria pertusa 25<br />

Pesotum 366<br />

Petriella 368<br />

Peziza 419<br />

Peziza succosa 419<br />

Peziza vesiculosa Pl. 6<br />

Pezizales (operculate discomycetes)<br />

414–428, Pl. 6; classification 415;<br />

phylogeny 415<br />

Pezoloma 690<br />

pH sensing 433<br />

Phaeolus 576<br />

Phaeosphaeria 460, 462–464;<br />

anamorphs 461<br />

Phaeosphaeria avenaria (anam.<br />

Stagonospora avenae) 462<br />

Phaeosphaeria (anam. Stagonospora)<br />

nodorum 231, 462–463, 481<br />

Phaffia rhodozyma 665–666;<br />

see Xanthophyllomyces dendrorhous<br />

phagocy<strong>to</strong>sis 40, 45, 55<br />

phalloidin 539, 540<br />

Phallus duplicatus 591<br />

Phallus impudicus 15, 500, 591, 592,<br />

Pl. 11; mycelial cord 15, 591; odours<br />

591<br />

Phallus indusiatus 589, 591–592


INDEX<br />

833<br />

Phallus ravenelii 591<br />

Phanerochaete chrysosporium 565;<br />

white-rot 530–532<br />

Phanerochaete velutina 570<br />

phaneroplasmodium 48, 51, Pl. 1<br />

Phellinus 573<br />

Phellinus igniarius 573, Pl. 10<br />

Phellinus noxius 573<br />

Phellinus pomaceus 573; chloromethane<br />

emission 532, 574<br />

Phellinus robustus 574<br />

Phellinus weirii 573<br />

phenylacetaldehyde 591, 630<br />

phenylamides 112<br />

phenylethanol 591, 630<br />

pheromones; see sex hormones<br />

phialide 31, 232, 233–234, 300<br />

phialoconidium 300, 465; dry and<br />

slimy 233<br />

Phialophora 386, 439<br />

Phlebiopsis gigantea 569<br />

phoenicoid fungi 417<br />

Pholiota 554<br />

Pholiota nameko 555<br />

Pholiota squarrosa 554, Pl. 9<br />

Phoma 462, 464–465, 466; versus<br />

Phomopsis 465<br />

Phoma acuta 462–463<br />

Phoma betae 466<br />

Phoma epicoccina 466; see Epicoccum<br />

nigrum<br />

Phoma lingam 463; Plenodomus 464;<br />

see Lep<strong>to</strong>sphaeria maculans<br />

Phoma medicaginis 464, 465<br />

phomopsins 375<br />

Phomopsis 373–375; a-conidium 373,<br />

374; b-conidium 373, 374, 464;<br />

ecology 375; versus Phoma 465;<br />

taxonomy 374<br />

Phomopsis helianthi 375<br />

Phomopsis phaseoli 373; see Diaporthe<br />

phaseolorum<br />

Phomopsis viticola 375<br />

phomozin 375<br />

phosphonate fungicides 113<br />

pho<strong>to</strong>biont (of lichens) 446<br />

pho<strong>to</strong>recep<strong>to</strong>rs 169<br />

pho<strong>to</strong>sensitizers 482, 665<br />

pho<strong>to</strong>tropism, ascus tip 420, 427;<br />

sporangiophore 168, 169, 170, 185,<br />

188<br />

Phragmidium 631<br />

Phragmidium mucronatum 631;<br />

teliospore 629<br />

Phragmidium violaceum 631, Pl. 12;<br />

biocontrol of bramble 631<br />

phragmobasidium 488, 489, 593,<br />

637<br />

phragmospore 23<br />

Phycomyces blakesleeanus 168, 169, 184;<br />

heterothallism 174; life cycle 180;<br />

pho<strong>to</strong>tropism 168, 169–170;<br />

sporangiophore development<br />

168–170, 171; sporangium<br />

development 171–173; zygospore<br />

development 176, 186; zygospore<br />

germination 178<br />

Phycomyces nitens 179, 184<br />

Phycomycetes 33<br />

Phyllactinia 405<br />

Phyllactinia guttata 237, 390, 406, 407;<br />

chasmothecium 407, 409<br />

phylloplane yeasts 668<br />

phylogeny 33, 36–37, 69, 166, 248,<br />

512, 516<br />

Physarales (Myxomycota) 51<br />

Physarum polycephalum 48, 50, 51–53,<br />

Pl. 1; life cycle 49; smart network 52<br />

physiological heterothallism 435<br />

physiological races; see races<br />

Physoderma alfalfae 153<br />

Physoderma maydis 153<br />

phy<strong>to</strong>alexins 115, 383, 438, 464<br />

Phy<strong>to</strong>phthora 95, 102–115; asexual<br />

reproduction 103–106;<br />

chlamydospore 105, 106; oospore<br />

109–110; sexual reproduction<br />

109–111; sporangium germination<br />

104; zoospore 106, 107–108;<br />

zoospore encystment 106; zoospore<br />

release 105<br />

Phy<strong>to</strong>phthora alni 103<br />

Phy<strong>to</strong>phthora cac<strong>to</strong>rum 102, 106,<br />

109–110<br />

Phy<strong>to</strong>phthora cinnamomi 102, 107,<br />

108<br />

Phy<strong>to</strong>phthora drechsleri 103<br />

Phy<strong>to</strong>phthora erythroseptica 29, 68, 103,<br />

109, 116, Pl. 2<br />

Phy<strong>to</strong>phthora fragariae 103<br />

Phy<strong>to</strong>phthora infestans 78, 95, 111–115;<br />

disease forecast 113; epidemiology<br />

111–112; fungicides 112–113;<br />

haus<strong>to</strong>rium 103, 105; life cycle<br />

95–96, 101, 102; origin and spread<br />

111; physiological races 114; pota<strong>to</strong><br />

late blight 111–115, Pl. 2; <strong>to</strong>ma<strong>to</strong><br />

late blight 115<br />

Phy<strong>to</strong>phthora nicotianae 103<br />

Phy<strong>to</strong>phthora palmivora 103<br />

Phy<strong>to</strong>phthora ramorum 103<br />

phy<strong>to</strong><strong>to</strong>xins 375, 388, 470, 476, 478,<br />

482, 568; see cercosporin, HC-<strong>to</strong>xin,<br />

prehelminthosporol, T-<strong>to</strong>xin,<br />

vic<strong>to</strong>rin C<br />

Pichia 281<br />

Pichia guilliermondii 263<br />

Pichia jadinii 281<br />

Pichia pas<strong>to</strong>ris 263, 281<br />

Pichia stipitis 262<br />

Piedmont truffle 426<br />

pigments; see carotenoids, lichens,<br />

melanin, pulvinic acids<br />

Pilaira 189–190<br />

Pilaira anomala 190<br />

pileocystidium 525<br />

pileus 517<br />

Pilobolus 185–189; sporangium<br />

discharge 189; vec<strong>to</strong>r for<br />

nema<strong>to</strong>des 188<br />

Pilobolus crystallinus 185, Pl. 3<br />

Pilobolus kleinii 185, 188<br />

Pilobolus umbonatus 185<br />

pine, needlecast 441; rusts 634<br />

pionnote 345<br />

Pip<strong>to</strong>cephalis 201; appressorium 202;<br />

haus<strong>to</strong>rium 202<br />

Pip<strong>to</strong>cephalis freseniana 201<br />

Pip<strong>to</strong>cephalis fimbriata 202<br />

Pip<strong>to</strong>cephalis unispora 201<br />

Pip<strong>to</strong>cephalis virginiana 201, 202<br />

Pip<strong>to</strong>cephalis xenophila 201<br />

Pip<strong>to</strong>porus betulinus 339, 560, 561,<br />

562–563; mating system 506<br />

Piromyces 151<br />

Pisolithus 12, 527, 586; bole<strong>to</strong>id clade<br />

579; ec<strong>to</strong>mycorrhiza 586<br />

Pisolithus marmoratus 586<br />

Pisolithus tinc<strong>to</strong>rius (¼ P. arhizus) 586,<br />

Pl. 11<br />

Placopsis gelida 451<br />

plant growth hormones; see auxins,<br />

cy<strong>to</strong>kinins, ethylene, gibberellins<br />

plant resistance; see host resistance,<br />

breeding for resistance<br />

plasmodiocarp (of Myxomycetes) 50<br />

Plasmodiophora brassicae 54–57, 58,<br />

59–61; club root symp<strong>to</strong>ms 55–56,<br />

57; club root control 62–63;<br />

cruciform mi<strong>to</strong>sis 55, 59; life cycle<br />

56; infection process 57, 58–59, 65;<br />

plant hormones 60–61, 253


834 INDEX<br />

Plasmodiophoromycota 24, 38, 54–66;<br />

amoebae 55; phagocy<strong>to</strong>sis 55;<br />

taxonomy 54<br />

plasmodium 3, 40, 47, 55;<br />

see amoeboid, aphano-, phanero-,<br />

pro<strong>to</strong>-<br />

Plasmopara nivea 119<br />

Plasmopara pusilla 120<br />

Plasmopara pygmaea 120<br />

Plasmopara viticola 78, 116, 119, 410<br />

plectenchyma 520<br />

Plec<strong>to</strong>mycetes 285–314, Pl. 4;<br />

lichenized ances<strong>to</strong>rs? 454;<br />

taxonomy 286<br />

Plenodomus 464<br />

pleomorphism 32<br />

Pleospora 460, 466–469, 472<br />

Pleospora bjoerlingii (¼ P. betae) 466<br />

Pleospora herbarum 31, 466, 468;<br />

ascospore perisporium 237, 468;<br />

pseudothecium development<br />

461<br />

Pleospora scirpicola 466; ascospore<br />

appendage 239, 460; bitunicate<br />

ascus 460<br />

Pleosporales 460–480; pseudothecium<br />

development 460–461<br />

pleurocystidium 525<br />

Pleurotus 541–542; cultivation 525,<br />

542; nema<strong>to</strong>de-trapping 679, 681;<br />

nema<strong>to</strong><strong>to</strong>xic substances 683<br />

Pleurotus cystidiosus 542<br />

Pleurotus ostreatus 533, 542<br />

Pleurotus tuberregium 542<br />

plum, pocket plum 251; silver leaf<br />

571; witches’ broom 251<br />

Pluteus 541<br />

Pluteus cervinus 541; trama 524<br />

Pneumocystis 250, 259–261<br />

Pneumocystis carinii 259<br />

Pneumocystis jirovecii 259<br />

pneumonia 259<br />

Pochonia chlamydosporia, as nema<strong>to</strong>de<br />

egg parasite 684; biological control<br />

685<br />

Podaxis 578<br />

Podaxis pistillaris 578<br />

podetium 457<br />

Podocrella (anam. Harposporium) 349,<br />

681<br />

Podosordaria tulasnei 15, 18<br />

Podosphaera 404–405<br />

Podosphaera clandestina 404, 405<br />

Podosphaera leucotricha 390, 405<br />

Podosphaera xanthii, resistance<br />

breeding 408<br />

Podospora 320–326; versus Schizothecium<br />

320; mating systems 322–324<br />

Podospora anserina 320–326; ascospore<br />

development 321, 322; mating type<br />

idiomorphs 325; post-meiotic<br />

mi<strong>to</strong>sis 323–324; senescence 326;<br />

vegetative incompatibility 325–326,<br />

377<br />

Podospora curvicolla 320<br />

Podospora curvula 320<br />

Podospora decipiens 320, 322<br />

Podospora pleiospora 320, 322<br />

pollen (as substrate) 73, 74, 140<br />

polyketides 304<br />

Polymyxa betae 54, 57, 59, 62<br />

Polymyxa graminis 54, 62<br />

polyplanetism 81, 85<br />

polyploidy 159<br />

polyporoid clade 560–566; phylogeny<br />

516–517, 560<br />

Polyporus brumalis 561<br />

Polyporus mylittae 18, 20, 501, 502<br />

Polyporus squamosus 561, 562, 564<br />

Polyporus tuberaster 501<br />

polyunsaturated fatty acids (PUFA) 74,<br />

167, 197<br />

poplar, leaf blister 251<br />

poroconidium 232, 468, 469, 470,<br />

472, 473<br />

post-harvest diseases 388, 432<br />

post-meiotic mi<strong>to</strong>sis, in Ascomycota<br />

323–324; in Basidiomycota 490<br />

pota<strong>to</strong>, black dot disease 388; black<br />

scurf 595; late blight 111, Pl. 2;<br />

mop-<strong>to</strong>p virus 62–63; pink rot 103,<br />

Pl. 2; powdery scab 61, 63; wart<br />

disease 134, Pl. 3<br />

powdery mildews 390; control by<br />

breeding, fungicides, biological<br />

control 408–413; see Erysiphales<br />

PR proteins 119, 410<br />

predacious yeasts 282, 283<br />

prehelminthosporol 475, 476<br />

pre-thallus (lichens) 450<br />

primary mycelium 492, 496<br />

principal host 610<br />

principal zoospore 76<br />

prion-like proteins 326<br />

probasidium 490, 610, 637<br />

proconidium 327<br />

progametangium 174, 176<br />

promycelium 488, 489, 611, 612, 637<br />

prosorus 134<br />

proteases 683<br />

protein glycosylation 11, 654<br />

protenchyma 519<br />

pro<strong>to</strong>-aecium 611, 622<br />

pro<strong>to</strong>corm 596<br />

Pro<strong>to</strong>myces 250–251, 261<br />

pro<strong>to</strong>n pumps; see H þ ATPases<br />

pro<strong>to</strong>perithecium 318, 329<br />

pro<strong>to</strong>plasmodium 48<br />

Pro<strong>to</strong>steliomycetes (pro<strong>to</strong>stelid<br />

plasmodial slime moulds)<br />

45–47<br />

Pro<strong>to</strong>stelium 46, 47<br />

pro<strong>to</strong>tunicate; see ascus wall<br />

Pro<strong>to</strong>zoa 37, 40, 54, 67<br />

Psathyrella 537<br />

Pseudeurotium, sec<strong>to</strong>ring 229<br />

Pseudoaegerita 701<br />

pseudoaethalium 50<br />

Pseudoallescheria 368<br />

Pseudocercospora 481<br />

Pseudocercospora herpotrichoides 439<br />

Pseudogymnoascus 444<br />

pseudohomothallism 322, 323, 330,<br />

506, 570; see secondary<br />

homothallism<br />

Pseudohydnum gelatinosum 604<br />

pseudohypha 2, 3, 227, 272<br />

Pseudoidium 401, 402<br />

Pseudomonas, in biological control 347,<br />

386, 596; induction of fruiting in<br />

Agaricus 534<br />

pseudoparaphysis 242, 460<br />

pseudoparenchyma 285, 520<br />

pseudoplasmodium 3, 42, 44<br />

pseudorhiza 547, 548<br />

pseudosclerotium 550<br />

pseudothecium 245, 246; Pleospora<br />

type 460–461<br />

Pseudotulos<strong>to</strong>ma 314<br />

Pseudozyma 412<br />

psilocin 539, 554<br />

Psilocybe 554<br />

Psilocybe cubense 554<br />

Psilocybe merdaria, secotioid form<br />

578<br />

Psilocybe mexicana 554<br />

Psilocybe semilanceata 553, 554<br />

psilocybin 539, 546, 554<br />

psoromic acid 453<br />

Puccinia, pathway 624; versus Uromyces<br />

629<br />

Puccinia caricina 630, Pl. 12


INDEX<br />

835<br />

Puccinia coronata 476, 612, 619, 628;<br />

homothallism 628; host resistance<br />

genes 628; teliospore 629;<br />

Tranzschel’s Law 614<br />

Puccinia distincta Pl. 12<br />

Puccinia graminis 620, 621–627; on<br />

barberry 621–622; basidiospores<br />

621; basidium 489; crop losses 620;<br />

cultivation on simple media 618;<br />

host Sr resistance genes 626;<br />

life cycle 610, 611, 612–613, 621–622;<br />

formae speciales 621; f. sp. secalis 622;<br />

f. sp. tritici 622, 624–627;<br />

physiological races 621, 625, 626,<br />

627; somatic hybridization 625<br />

Puccinia hordei 628<br />

Puccinia lagenophorae 612, 613<br />

Puccinia melanocephala 633<br />

Puccinia menthae 613, 630<br />

Puccinia mesnieriana 614<br />

Puccinia monoica 630; floral mimicry<br />

630<br />

Puccinia obscura 612<br />

Puccinia poarum 630<br />

Puccinia punctiformis 630<br />

Puccinia recondita 627<br />

Puccinia sorghi 629<br />

Puccinia striiformis 408, 627–628<br />

Puccinia triticina 627; host Lr resistance<br />

genes 627<br />

PUFA; see polyunsaturated fatty acids<br />

puff balls 581, 582, Pl. 11<br />

puffing; see ascospore puffing<br />

pullulan 485<br />

pulvinic acids 453, 455, 555, 557, 585<br />

pycnidiospore 465<br />

pycnidium 231, 374, 462, 464,<br />

465–466<br />

Pycnoporus cinnabarinus 531<br />

pyramiding (of resistance alleles) 408,<br />

619, 627<br />

Pyrenomycetes 315, 316–366,<br />

367–389, Pl. 5; taxonomy 316<br />

Pyrenopeziza brassicae 436, 439, 440<br />

Pyrenophora (anam. Drechslera) 460,<br />

472, 477–478; phy<strong>to</strong><strong>to</strong>xins 478<br />

Pyrenophora teres 478<br />

Pyrenophora tritici-repentis 476, 478<br />

Pyricularia grisea 378<br />

Pyricularia oryzae (teleom. Magnaporthe<br />

grisea) 378<br />

Pyronema 415–417; antheridium 416;<br />

ascogenous hypha 416; ascogonium<br />

416; crozier 416; ecology 417;<br />

investing hyphae 416; trichogyne<br />

416<br />

Pyronema domesticum 229, 415, 416,<br />

417, 418, Pl. 6; sclerotium 415<br />

Pyronema omphalodes 415, 417<br />

Pythiales 95–115; asexual<br />

reproduction 95; hormones 96;<br />

sexual reproduction 95, 100<br />

Pythioge<strong>to</strong>n zeae 95<br />

Pythiopsis cymosa 92, 93<br />

Pythium 95–96, 97, 98–102, Pl. 2;<br />

asexual reproduction 97–100;<br />

biological control 340; ecology<br />

102; sexual reproduction<br />

100–102<br />

Pythium acanthicum 97<br />

Pythium aphanidermatum 97–98<br />

Pythium debaryanum 97, 98, 100, 101<br />

Pythium gracile 97<br />

Pythium heterothallicum 101<br />

Pythium insidiosum 79, 102<br />

Pythium mamillatum 100, 101<br />

Pythium middle<strong>to</strong>nii 99<br />

Pythium monospermum 69<br />

Pythium multisporum 95<br />

Pythium nunn 97<br />

Pythium oligandrum 97<br />

Pythium splendens 101<br />

Pythium sylvaticum 95, 101<br />

Pythium ultimum 97, 100, 101<br />

Pythium undulatum 30, 99, Pl. 2<br />

quaterverticillate; see Penicillium<br />

conidiophores<br />

quinoxyfen 410, 412<br />

Quorn 339<br />

races 114, 138, 395, 619, 621, 643<br />

radioactive pollution 454<br />

Ramaria 576<br />

Ramaria botrytis 576, Pl. 10<br />

Ramaria stricta 576<br />

Ramichloridium 486<br />

ramoconidium 483<br />

ramus 310, 311<br />

RAPD analysis 34<br />

raspberry, cane blight 461<br />

reactive oxygen species (ROS) 326,<br />

482, 665<br />

receptacle 591, 592<br />

receptive hypha 612<br />

recombinant proteins, production<br />

263, 281<br />

red yeasts; see basidiomycete yeasts<br />

regulated secretion (in zoospore<br />

encystment) 77, 106<br />

reindeer lichen 458<br />

relative sexuality 77, 91, 95<br />

replacement disease 350<br />

resistance; see breeding, fungicide<br />

resistance, host resistance<br />

respira<strong>to</strong>ry allergens; see asthma<br />

resting sporangium 56, 144, 147, 149,<br />

157, 159, 161<br />

Reticularia lycoperdon 50, 51, Pl. 1<br />

reticuloperidium 285, 289, 296–297<br />

retraction septum 203, 602<br />

RFLP analysis 34<br />

rhexolytic secession 235, 285, 289, 505<br />

Rhizina undulata 414<br />

rhizinae 447, 456, 457, Pl. 8<br />

Rhizocarpon geographicum 447, 451, Pl. 8<br />

Rhizoc<strong>to</strong>nia 594–598; anas<strong>to</strong>mosis<br />

groups 594; biological control 340,<br />

596; hyphae 594, 597; sclerotia 18,<br />

19, 501<br />

Rhizoc<strong>to</strong>nia cerealis (teleom.<br />

Cera<strong>to</strong>basidium cornigerum) 596;<br />

orchid mycorrhiza 596, 597<br />

Rhizoc<strong>to</strong>nia goodyerae-repentis 596<br />

Rhizoc<strong>to</strong>nia repens 596<br />

Rhizoc<strong>to</strong>nia solani (teleom.<br />

Thanatephorus cucumeris) 594–595;<br />

anas<strong>to</strong>mosis groups 594; orchid<br />

mycorrhiza 596; sclerotium 595<br />

rhizoids 3, 128, 141, 149–150, 155<br />

rhizomorph 16, 17–18, 500, 502, 550<br />

rhizomycelium 128, 142, 143, 144, 145<br />

Rhizophlyctis 148<br />

Rhizophlyctis harderi 131<br />

Rhizophlyctis oceanis 150<br />

Rhizophlyctis rosea 129, 147–148,<br />

149–150, 153; cellulose degradation<br />

148<br />

Rhizophydium 133, 139, 140, 141–142;<br />

fossil 133<br />

Rhizophydium plank<strong>to</strong>nicum 140–142;<br />

parasitizing Asterionella 141<br />

Rhizophydium pollinis-pini 139<br />

Rhizophydium sphaerocarpon 139<br />

Rhizopogon 558, 587, Pl. 11; bole<strong>to</strong>id<br />

clade 579<br />

Rhizopogon luteolus 588<br />

Rhizopogon roseolus 588<br />

Rhizopogon vinicolor 588<br />

Rhizopus 182–183<br />

Rhizopus arrhizus 182<br />

Rhizopus azygosporus 179


836 INDEX<br />

Rhizopus microsporus 182<br />

Rhizopus oryzae 182<br />

Rhizopus rhizopodiformis 182<br />

Rhizopus sexualis 29, 173, 177, 183;<br />

zygospore development 176, 178<br />

Rhizopus s<strong>to</strong>lonifer 166, 172–173, 176,<br />

179, 182, 184, Pl. 3<br />

rhizoxin 183<br />

Rhodocollybia 546<br />

Rhodosporidium <strong>to</strong>ruloides 667<br />

Rhodo<strong>to</strong>rula 667; carotenoid<br />

production 671; ecology 668<br />

Rhodo<strong>to</strong>rula glutinis (teleom.<br />

Rhodosporidium) 659, 667<br />

Rhynchosporium secalis 439<br />

Rhynie chert 35, 133, 246, 454<br />

Rhytisma acerinum 441, 442–443<br />

Rhytismataceae 440<br />

riboflavin production 284<br />

rice, foolish seedling disease 339;<br />

rice blast 378<br />

ring wall building 232<br />

ringworm 293<br />

Robigalia 610<br />

roestelioid aecium 632, Pl. 12<br />

Rogation Sunday 610<br />

root exudates 434, 435<br />

root rot, Phy<strong>to</strong>phthora and Pythium 102,<br />

Pl. 2<br />

rose, powdery mildew 404; rust 631<br />

Rosellinia necatrix 333<br />

rostrate; see ascus wall<br />

rotifers 64<br />

rRNA (ribosomal RNA) analysis 35, 36<br />

rumen fungi; see Neocallimastigales<br />

rumposome 130<br />

Russula 566; ec<strong>to</strong>mycorrhiza 566;<br />

orchid mycorrhiza 598<br />

Russula atropurpurea 567<br />

Russula cyanoxantha, trama 524<br />

Russula emetica 566<br />

Russula fellea 566<br />

Russula ochroleuca 566<br />

russuloid clade 566–572; phylogeny<br />

516–517<br />

rust fungi 610; cereal rusts 627–629;<br />

see Uredinales<br />

Rutstroemia echinophila Pl. 7<br />

rye, ergot 350; eyespot 439<br />

Saccharomyces 263<br />

Saccharomyces cerevisiae 10–11, 229,<br />

232, 256; alcoholic fermentation<br />

262, 274–275; ascospore cleavage<br />

267; bread-making 274; cell cycle<br />

270, 271, 272, 301; cell wall 270;<br />

cy<strong>to</strong>logy 264; killer <strong>to</strong>xins 273–274,<br />

646; life cycle 265–266; mating<br />

266–270; membrane cycling 272;<br />

morphogenesis 270–272;<br />

pheromones 267; pseudohyphae<br />

272; vacuole 273<br />

Saccharomyces pas<strong>to</strong>rianus 263<br />

Saccharomycopsis 282–284<br />

Saccharomycopsis fibuligera 282, 283<br />

Saccobolus 420<br />

sagenogen 68, 73<br />

Sai<strong>to</strong>ella 250<br />

Sakaguchia dacryoidea 659<br />

saké 276<br />

salicylic acid 412<br />

sanguinolen<strong>to</strong>us hypha; see lactifer<br />

Saprochaete 282<br />

Saprolegnia 9, 24, 30, 79, 81–86,<br />

Pl. 2; asexual reproduction 82,<br />

83–85; cyst 84; hypha 80; life cycle<br />

78; ooplast arrangement 86;<br />

sexual reproduction 85, 86;<br />

zoospore 76<br />

Saprolegnia li<strong>to</strong>ralis 85<br />

Saprolegnia parasitica 81, 85<br />

Saprolegnia polymorpha 81<br />

Saprolegniales 79–95; hypha 79;<br />

nutrition 81<br />

sap-stain (of wood) 364, 371, 373<br />

sarco-hypha 518<br />

sarcomitic basidiocarp construction<br />

519<br />

sarcotrimitic basidiocarp<br />

construction 519<br />

Sarcoscypha 419<br />

Sarcoscypha australis Pl. 6<br />

Sauternes wine 438<br />

Sawadaea 403<br />

Sawadaea bicornis 404, 405<br />

Schizochytrium 74<br />

schizogenous perithecial<br />

development 315, 318, 338<br />

schizolytic secession 235<br />

Schizophyllum 542<br />

Schizophyllum commune 7–8, 489, 499,<br />

542, 543–544, 545; dikaryotization<br />

(Buller phenomenon) 510; fruiting<br />

in the labora<strong>to</strong>ry 544;<br />

hydrophobins 545; mating type<br />

fac<strong>to</strong>rs 508, 543<br />

Schizosaccharomyces 250, 261<br />

Schizosaccharomyces japonicus 253<br />

Schizosaccharomyces oc<strong>to</strong>sporus 253,<br />

254, 255<br />

Schizosaccharomyces pombe 51, 253–259;<br />

cell cycle 256, 257, 301; cell walls<br />

255; life cycle 255; morphogenesis<br />

258, 259; pheromones 255<br />

Schizosaccharomycetales 253–259<br />

Schizothecium 320<br />

Schizothecium tetrasporum 321, 322<br />

Schizothecium vesticola 320<br />

Scleroderma 557, 585–586; bole<strong>to</strong>id<br />

clade 579, 585<br />

Scleroderma bovista 586<br />

Scleroderma citrinum 515, 586, 587<br />

Scleroderma verrucosum 586, 587<br />

Sclerosporaceae 125–126<br />

Sclerospora 125<br />

Sclerospora graminicola 126<br />

sclerotial stroma 430<br />

Sclerotinia, macroconidium 430;<br />

microconidium 430, 432;<br />

sclerotium (sclerotial stroma) 20,<br />

430<br />

Sclerotinia (Myriosclerotinia) curreyana<br />

430, 431, Pl. 7; Myrioconium-type<br />

microconidium 430, 431;<br />

sclerotium 431<br />

Sclerotinia fructigena; see Monilinia<br />

fructigena<br />

Sclerotinia fuckeliana; see Botryotinia<br />

fuckeliana<br />

Sclerotinia laxa; see Monilinia laxa<br />

Sclerotinia (Stromatinia) narcissi, mating<br />

behaviour 435<br />

Sclerotinia porri; see Botryotinia porri<br />

Sclerotinia sclerotiorum 431–434;<br />

biological control 433; infection<br />

biology 432; mating behaviour 436;<br />

oxalic acid 433; pH sensing 433;<br />

sclerotium 432, 434<br />

Sclerotinia trifoliorum, mating<br />

behaviour 436<br />

Sclerotinia (Dumontinia) tuberosa 431,<br />

432<br />

sclerotium 18, 19, 20–21; in<br />

Myxomycetes 49; see Aspergillus,<br />

Botrytis, Claviceps, Cordyceps,<br />

Morchella, Penicillium, Polyporus<br />

mylittae, Rhizoc<strong>to</strong>nia, Sclerotinia,<br />

Sclerotium<br />

Sclerotium cepivorum 431, 434; sclerotia<br />

and root exudates 434, 435<br />

Sclerotium rolfsii, sclerotium 20, 501<br />

scolecospore 23


INDEX<br />

837<br />

scolytid beetles 366<br />

Scopulariopsis 368<br />

Scopulariopsis brevicaulis 234, 369<br />

Scytalidium thermophilum 534<br />

secondary metabolism 453<br />

secondary mycelium 492, 497<br />

secondary resource capture 554<br />

secondary sporidium 648, 649<br />

secotioid fruit-bodies 578<br />

secretion of proteins 10–11, 13<br />

sec<strong>to</strong>ring (of mycelium) 228, 229<br />

seed-borne fungi 433, 463, 466, 470,<br />

478, 643<br />

seed certification 651<br />

seed dressing 651<br />

segregation, first-division versus<br />

second-division in meiosis 319<br />

senescence 326<br />

separating cell 235<br />

Sepedonium chrysospermum (teleom.<br />

Apiocrea chrysosperma) 556, 585, Pl. 9<br />

septins 271, 301<br />

Sep<strong>to</strong>ria 462, 481<br />

Sep<strong>to</strong>ria tritici 481, 483<br />

septum, Ascomycota 227, 228;<br />

dis<strong>to</strong>septate versus euseptate 472;<br />

formation 258, 259; in hypha 2;<br />

pores 261; see dis<strong>to</strong>septate,<br />

dolipore, parenthesome<br />

Serpula lacrymans 556, 558–560;<br />

control 560; mycelial cord 16, 559;<br />

reproduction 559; var. domesticus<br />

559–560<br />

sex chromosomes 638, 653, 663<br />

sex hormones (pheromones);<br />

see Achlya, Allomyces, Ascobolus<br />

furfuraceus, Blakeslea trispora,<br />

Saccharomyces cerevisiae,<br />

Schizosaccharomyces pombe, Tremella<br />

mesenterica, Ustilago maydis<br />

sexual agglutination;<br />

see agglutination<br />

sexual dimorphism 436<br />

shiitake mushroom; see Lentinula<br />

edodes<br />

shuttle-streaming 51<br />

side body complex 155<br />

signalling 10, 277, 384, 645; in<br />

appressorium formation 384–385,<br />

396; calcineurin 664; cross-talk 385,<br />

646; heterodimer formation 509,<br />

645; hydrophobicity 384;<br />

mycoviruses 376; ring constriction<br />

in Drechslerella 678<br />

Simulium, as hosts for Erynia conica<br />

213; and trichomycetes 223–224<br />

single-cell protein (SCP) 263<br />

sinuate gills 523<br />

sirenin 158<br />

Sis<strong>to</strong>trema hamatum (anam. Ingoldiella<br />

hamata) 504, 689<br />

skeletal hypha 518<br />

skele<strong>to</strong>-ligative hypha 518<br />

slime moulds 38, 40, Pl. 1<br />

slime moulds (cellular); see Acrasiomycetes,<br />

Dictyosteliomycetes<br />

slime moulds (plasmodial); see<br />

Pro<strong>to</strong>steliomycetes, Myxomycetes<br />

slime net 71, 72, 73, 74<br />

Smittium culisetae 222, 224<br />

smut fungi 636; covered smut 643;<br />

loose smut 642, 643;<br />

see Microbotryales, ustilaginomyce<strong>to</strong>us<br />

smuts<br />

soft-rot (of wood) 331<br />

somatic diploids<br />

somatic hybridization 625<br />

soma<strong>to</strong>gamy 132, 229<br />

soralium 448<br />

Sordaria 317, 318–320; ascospore<br />

discharge 244; mating systems<br />

318–319<br />

Sordaria brevicollis 320<br />

Sordaria fimicola 68, 229; ascospore<br />

perisporium 237; ascus apical<br />

apparatus 241<br />

Sordaria heterothallis 320<br />

Sordaria humana, ascospore wall 237,<br />

239<br />

Sordaria macrospora 320<br />

Sordariales 315, 316, 317–332<br />

soredium 448, 450<br />

sorghum, ergot 349; anthracnose 388<br />

sorocarps 41, 45<br />

sorus 60, 636, 639<br />

sour rot 282<br />

soy sauce 302<br />

Sparassis 564<br />

Sparassis crispa 561<br />

spectrin 9<br />

spermatium 231, 329, 376, 431, 435,<br />

442, 611, 622<br />

sperma<strong>to</strong>zoid 162, 164<br />

spermodochidium 431<br />

spermogonium 442, 611, 612, 622,<br />

Pl. 12<br />

Sphacelia segetum (teleom. Claviceps<br />

purpurea) 349<br />

Sphaerobolus 588; gomphoid-phalloid<br />

clade 579<br />

Sphaerobolus stellatus 588, 589–590;<br />

peridiole 589–590; peridiole<br />

discharge 589<br />

sphaerocyst 519<br />

Sphaeronaemella fimicola 371, 372<br />

Sphaeronaemella helvellae 371<br />

Sphaerotheca 404–405; fibrosin bodies<br />

404<br />

Sphaerotheca (Podosphaera) fuliginea 404<br />

Sphaerotheca (Podosphaera) macularis<br />

404; resistance breeding 408<br />

Sphaerotheca (Podosphaera) mors-uvae<br />

390, 404; resistance breeding 408<br />

Sphaerotheca (Podosphaera) pannosa 404,<br />

405–406<br />

spherule, of Myxomycetes 49; in<br />

Coccidioides 291–292<br />

Spinellus fusiger 166, 551, Pl. 3<br />

Spiniger 504, 568; see Heterobasidion<br />

annosum<br />

spiral growth 171<br />

Spirosphaera 697, 700<br />

Spirosphaera carici-graminis 701<br />

Spirosphaera floriformis 701<br />

Spirosphaera minuta 701<br />

Spitzenkörper; see hypha<br />

Spizellomyces 145<br />

Spizellomycetales 145–150<br />

splash cup 580, 582, 584–585<br />

Spongospora nasturtii 54, 59, 61<br />

Spongospora subterranea 54–55, 59,<br />

61–62; vec<strong>to</strong>r for plant viruses 62<br />

sporabola 495<br />

sporangiolum 166, 191, 193, 194,<br />

196, 197<br />

sporangiophore, development<br />

168–170, 171; microphores and<br />

macrophores 168; in Mucorales 165<br />

sporangiospore 24, 165, 168<br />

sporangium 25; one-spored 32;<br />

see sporangiolum; of Mucorales<br />

171–173; of Myxomycetes 50; of<br />

Phy<strong>to</strong>phthora 104, 106; of Pythium<br />

98–99; of Saprolegnia 82<br />

spore print 532<br />

spores, different kinds 22–23; see also<br />

conidium, numbers of amero-,<br />

aplano-, asco-, ballis<strong>to</strong>-, basidio-,<br />

chlamydo-, dictyo-, didymo-, helico-,<br />

memno-, oo-, phragmo-, scoleco-,<br />

sporangio-, statismo-, stauro-, xeno-,<br />

zoo-, zygo-


838 INDEX<br />

Sporidiales 609, 666–670;<br />

see urediniomycete yeasts<br />

Sporidiobolus 659; see Sporobolomyces<br />

Sporidiobolus ruineniae 659<br />

Sporidiobolus salmonicolor 667, 669;<br />

life cycle 670<br />

sporidium, in Hap<strong>to</strong>glossa 64, 65, 66;<br />

in smut fungi 637, 648<br />

Sporisorium 640<br />

Sporisorium scitamineum 639<br />

Sporobolomyces 658–659; ecology 668<br />

Sporobolomyces roseus (teleom.<br />

Sporidiobolus) 659, 666–667, 668; air<br />

pollution 669; carotenoid<br />

production 671; ecology 668<br />

Sporobolomyces salmonicolor (teleom.<br />

Sporidiobolus) 667, 670<br />

sporocarp 46, 47, Pl. 1<br />

sporocyst 288<br />

sporodochium 231, 342–343, 466, Pl. 5<br />

sporophores, of Myxomycetes 50<br />

sporopollenin 177<br />

Sporormiella 460–461, 479<br />

Sporormiella intermedia 479, 480<br />

Sporoschisma 234<br />

Sporothrix 366<br />

Sporothrix schenckii 364<br />

Sporotrichum 505, 564<br />

Sporotrichum pulverulentum 566<br />

sprinkle-plate technique 675<br />

squamule 448<br />

squamulose lichens 447<br />

Stagonospora 462, 464<br />

Stagonospora avenae; see Phaeosphaeria<br />

avenaria<br />

Stagonospora nodorum; see Phaeosphaeria<br />

nodorum<br />

St. Anthony’s fire; see ergotism<br />

statismospore 28, 577<br />

sta<strong>to</strong>lith 548<br />

staurospore 23<br />

steliogen 46<br />

Stemonitis axifera 51, Pl. 1<br />

Stemphylium (teleom. Pleospora) 467,<br />

468, 472<br />

Stemphylium vesicarium 468<br />

stereothecium 423<br />

Stereum 569–572, 605; mating 569<br />

Stereum gausapatum 569<br />

Stereum hirsutum 567, 569–570; bow-tie<br />

reaction 571; mating 569<br />

Stereum rugosum 568, 569<br />

Stereum sanguinolentum 569; mating<br />

570<br />

sterigma 489, 494<br />

sterigma<strong>to</strong>cystin 305<br />

sterols 89, 96, 424; biosynthesis 411<br />

stichobasidium 490<br />

stinkhorns 589, 590<br />

stipe 517<br />

s<strong>to</strong>rage rots; see post-harvest diseases<br />

Straminipila 37, 67, 75; phylogeny 69<br />

straminipilous flagellum; see flagella<br />

strobilurins 410, 546, 551<br />

strobilurin-type fungicides 113, 410,<br />

412, 479, 481, 627, 631; resistance<br />

412<br />

Strobilurus 546<br />

Strobilurus tenacellus 412<br />

stroma 231, 430; see sclerotial stroma,<br />

substratal stroma<br />

Stropharia 554<br />

Stropharia aeruginosa 554<br />

Stropharia aurantiaca 554<br />

Stropharia semiglobata 521, 553, 554<br />

Stüben bodies 155<br />

stylospore 197<br />

sub-gleba 581, 583<br />

subs<strong>to</strong>matal vesicle 615, 616<br />

substratal stroma 430<br />

Subulicystidium longisporum 697<br />

sudden infant death syndrome 259<br />

sugar beet, blackleg 466; leaf spot<br />

481; powdery mildew 402;<br />

yellow-vein virus 62<br />

sugarcane, rust 633; smut 639<br />

Suillus 558; related <strong>to</strong> gasteromycetes<br />

578–579<br />

Suillus bovinus 558<br />

Suillus granulatus 558, Pl. 9<br />

Suillus grevillei 526, 556, 558; related <strong>to</strong><br />

gasteromycetes 578<br />

Suillus luteus 558<br />

sulphur dioxide 454<br />

sulphur dust and lime 403, 410<br />

suprahilar plage 493<br />

surface tension catapult 493, 494, 496,<br />

611, 648, 655, 660<br />

survival of propagules 23, 51, 73, 111,<br />

155, 613, 643, 648<br />

suspensor 176, 177; appendages 177<br />

swarmer 46, 49, 137, 156, 160<br />

sympodula 332, 337, 472, 475<br />

synanamorph 32, 341–342, 366, 466<br />

Syncephalastrum racemosum 195<br />

Syncephalis 201<br />

Synchytrium 134–139<br />

Synchytrium aecidioides 139<br />

Synchytrium lagenariae 138<br />

Synchytrium endobioticum 127, 134–138,<br />

Pl. 3; control 138; hypertrophy 136;<br />

life-cycle 135<br />

Synchytrium fulgens 138<br />

Synchytrium macrosporum 134<br />

Synchytrium mercurialis 139, 140<br />

Synchytrium taraxaci 139, Pl. 3<br />

Synchytrium trichosanthidis 138<br />

synnema 230<br />

systemic acquired resistance (SAR)<br />

119, 412<br />

Syzygites megalocarpus 172–173,<br />

175–176, 179, 184, 187<br />

T-2 <strong>to</strong>xin 348<br />

Taeniomyces gracilis (teleom.<br />

Fibulomyces) 689, 690<br />

take-all 385–386; take-all decline 386,<br />

596<br />

Talaromyces 312, 313<br />

Tapesia acuformis 439–440, 441<br />

Tapesia yallundae (anam.<br />

Pseudocercospora herpotrichoides)<br />

439–440, 441, 596<br />

Taphrina 250–253, 261<br />

Taphrina amen<strong>to</strong>rum 251, Pl. 4<br />

Taphrina betulina 251, Pl. 4<br />

Taphrina epiphylla 253<br />

Taphrina deformans 251–252, 253, Pl. 4;<br />

plant hormones 253; control 253<br />

Taphrina insititiae 251<br />

Taphrina populina 251<br />

Taphrina pruni 251<br />

Taphrina <strong>to</strong>squinetii 251<br />

Taphrina wiesneri 251<br />

Taphrinales 251–253<br />

taxonomy 32–35, 38<br />

teleomorph 32<br />

teliospore 488, 489; pedicel 629, 631;<br />

in rust fungi 610, 612, 623; in smut<br />

fungi 637, 639, 660; in yeasts 667,<br />

670<br />

telium 610, 612, 623, Pl. 12<br />

terbinafine 279, 293<br />

Terfezia (desert truffle) 427<br />

Termi<strong>to</strong>myces 552<br />

terverticillate; see Penicillium<br />

conidiophores<br />

Tetrachaetum elegans 688, 689<br />

tetrapolar; see heterothallism<br />

tetraradiate propagules 214, 687–688,<br />

693<br />

textura angularis 464


INDEX<br />

839<br />

thallic conidiogenesis 30, 235;<br />

see holothallic, thallic-arthric<br />

thallic-arthric conidiogenesis 235<br />

thallus 3, 64, 73; epibiotic and<br />

endobiotic 128; holocarpic and<br />

eucarpic 71, 128; lichens 447–448,<br />

Pl. 8; monocentric and polycentric<br />

71, 128; variability 134;<br />

trichomycetes 223<br />

Thamnidium elegans 191, 192<br />

Thanatephorus 594, 595; see Rhizoc<strong>to</strong>nia<br />

Thaxterogaster 578<br />

Thelephora terrestris 572, Pl. 10; orchid<br />

mycorrhiza 598<br />

thelephoroid clade 572; phylogeny<br />

516–517<br />

thermo<strong>to</strong>lerance 306–307, 331<br />

Thielaviopsis 371<br />

Thielaviopsis basicola 232, 234, 372<br />

Thielaviopsis thielavioides 371<br />

thigmotropism 174–175, 615<br />

Thraus<strong>to</strong>chytriales 73–74<br />

Thraus<strong>to</strong>chytrium 73–74<br />

Thraus<strong>to</strong>theca clavata 91, 92<br />

thrush 279<br />

Tilletia 647; secondary sporidium 648,<br />

649; systemic infection 648<br />

Tilletia caries 647–648; teliospores 647;<br />

teliospore germination 648;<br />

trimethylamine 647<br />

Tilletia controversa 648<br />

Tilletia indica 649<br />

Tilletiopsis 412<br />

<strong>to</strong>bacco, blue mould 116, 119;<br />

necrosis virus (TNV) 148<br />

Tolypocladium inflatum 349, 363, 364;<br />

see cyclosporin<br />

<strong>to</strong>ma<strong>to</strong>, anthracnose 388; Fusarium<br />

wilt 347; late blight 115; leaf mould<br />

483<br />

Totivirus 273<br />

trama 524<br />

Trametes hirsuta 529<br />

Trametes (Coriolus) versicolor 499,<br />

560–562, 563, Pl. 10; vegetative<br />

incompatibility 510, 511, 518<br />

transgenic plants 433<br />

transposable elements 345, 439, 653<br />

Tranzschel’s Law 614<br />

Trebouxia (lichen pho<strong>to</strong>biont) 446,<br />

448, 450, 452, 455<br />

Tremella 605–608; basidiospore<br />

germination patterns 605, 608;<br />

haus<strong>to</strong>rial branches 605, 606<br />

Tremella encephala 605, 607<br />

Tremella foliacea 607<br />

Tremella frondosa 607, 608<br />

Tremella fuciformis 607;<br />

exopolysaccharides 608<br />

Tremella globospora 605, 606<br />

Tremella mesenterica 605, Pl. 11;<br />

life cycle 607; mating 605; sex<br />

hormones 605<br />

Tremellales 604–608; parenthesome<br />

604, 606; yeast stage 604<br />

tremelloid basidium 604, 606, 608<br />

tremerogen 605<br />

Trentepohlia (lichen pho<strong>to</strong>biont) 446,<br />

452, 455<br />

tretic conidial development 31,<br />

232, 472<br />

triadimefon 410, 411<br />

triazole-type fungicides 293, 306, 410,<br />

411, 627; resistance 411<br />

Trichia floriforme 51, 52<br />

Trichiales (Myxomycota) 51<br />

Trichoderma 340; biological control<br />

342, 434, 551, 569, 596<br />

‘Trichoderma effect’ (on sexual<br />

reproduction in Phy<strong>to</strong>phthora) 96,<br />

124<br />

Trichoderma harzianum 340, 342, 569<br />

Trichoderma reesii (teleom. Hypocrea<br />

jecorina) 339; cellulases 340<br />

Trichoderma stromaticum 551<br />

Trichoderma virens (teleom. Hypocrea<br />

virens) 340, 342<br />

Trichoderma viride (teleom. Hypocrea<br />

rufa) 340, 341<br />

Trichoglossum 443<br />

Trichoglossum hirsutum 443, 444<br />

trichogyne 330, 416<br />

Tricholoma 552<br />

Tricholoma gambosum 552<br />

Tricholoma matsutake 552; cultivation<br />

527<br />

Tricholoma sulphureum 552<br />

Trichomycetes 166, 222–225<br />

Trichophy<strong>to</strong>n 293, 294–295<br />

Trichophy<strong>to</strong>n mentagrophytes 294<br />

Trichophy<strong>to</strong>n rubrum 294<br />

Trichophy<strong>to</strong>n schoenleinii 294<br />

Trichophy<strong>to</strong>n verrucosum 294, 295<br />

trichospore 223, 224<br />

Trichosporon 660<br />

trichothecenes; see T-2 <strong>to</strong>xin<br />

Trichurus 371<br />

Tricladiomyces malaysianum 689<br />

Tricladium splendens (teleom.<br />

Hymenoscyphus splendens) 232,<br />

685, 688<br />

tridemorph 410, 411<br />

trimethylamine 647, 649<br />

trimitic basidiocarp construction 518,<br />

519<br />

Trimorphomyces papilionaceus 594<br />

tripartite tubular hair (TTH) 24,<br />

68–69, 70<br />

Triphragmium ulmariae 629<br />

trisporic acid 174, 175<br />

trophocyst 185, 187<br />

truffles 414; see Elaphomyces, Endogone,<br />

Rhizopogon, Tuber, Terfezia<br />

truffle flies 424<br />

TTH; see tripartite tubular hair<br />

T-<strong>to</strong>xin 476, 477<br />

Tuber 423–427; ascospore 425; boar<br />

hormone 5a-androst-16-en-3a-ol<br />

424; dimethyl sulphide 424;<br />

ec<strong>to</strong>mycorrhiza 424, 426; life cycle<br />

424–426; saprotrophic phase 426;<br />

yields 427<br />

Tuber aestivum 426<br />

Tuber albidum 426<br />

Tuber magnatum (Piedmont truffle) 426<br />

Tuber melanosporum (Périgord truffle)<br />

424–425, 426, 427; brûlé 426;<br />

inoculation of trees 427<br />

Tuber puberulum 425<br />

Tuber rufum 425<br />

Tubercularia state of Nectria cinnabarina<br />

231, 342, 344<br />

Tulasnellales 594; basidium 595<br />

Tulasnella 574<br />

Tumularia aquatica (teleom. Massarina<br />

aquatica) 685, 691, 693<br />

turbinate cells 142, 143<br />

turf-grass blight 72<br />

turgor pressure 7–8, 80, 83, 208; in<br />

ascospore discharge 242; in<br />

appressorium 381<br />

Typhula 501<br />

ubiquitin 272<br />

Umbelopsis 197<br />

Uncinocarpus 292<br />

Uncinula 401, 403<br />

Uncinula (Erysiphe) neca<strong>to</strong>r 390, 403,<br />

410<br />

uniseriate; see Aspergillus<br />

unitunicate; see ascus wall<br />

universal veil 520


840 INDEX<br />

Uredinales 609–635, Pl. 12; biotrophy<br />

616–618; control 627; cultivation<br />

on simple media 618; life cycle 610,<br />

611, 612–614; haus<strong>to</strong>rium 615–616,<br />

617–618; host resistance 618–620;<br />

infection process 614–616;<br />

monokaryotic versus dikaryotic<br />

penetration and haus<strong>to</strong>ria 614, 616;<br />

phylogeny 610; spore types<br />

611–613; systemic infections 630<br />

urediniomycete yeasts 659, 666–670;<br />

life cycle 667–668, 670;<br />

phylogeny 661, 666<br />

Urediniomycetes 609, Pl. 12;<br />

see Uredinales<br />

urediniospore 610, 612, 615, 623;<br />

adhesion pad 615; attachment 615<br />

uredinium 610, 612, 623, Pl. 12<br />

Urocystis 649–651<br />

Urocystis agropyri 649, 650<br />

Urocystis anemones 651; teliospore<br />

germination 650<br />

Urocystis tritici 649<br />

Uromyces 628, 630–631; infection<br />

processes monokaryotic versus<br />

dikaryotic 616<br />

Uromyces appendiculatus 614–615, 631;<br />

adhesion pad 615; teliospore 629;<br />

thigmotropism 615, 616<br />

Uromyces dactylidis 630<br />

Uromyces dianthi 630<br />

Uromyces ficariae 630<br />

Uromyces pisi 631<br />

Uromyces viciae-fabae 615–616, 631;<br />

haus<strong>to</strong>rium 617<br />

Usnea florida Pl. 8<br />

usnic acid 452, 453, 455, 458<br />

ustilaginomycete smuts 636–652;<br />

control 651–652; intracellular<br />

hyphae 637; life-cycle 637, 638;<br />

mating system 637; mating-type<br />

loci 637; sheath 637, 642; systemic<br />

growth 643; teliospore 637, 639<br />

ustilaginomycete yeasts 670–672;<br />

see Malassezia<br />

Ustilaginomycetes 636; subclasses<br />

636; see Exobasidiales,<br />

Malasseziales, ustilaginomyce<strong>to</strong>us<br />

smuts<br />

Ustilago 639–647; phylogeny 639;<br />

teliospore germination 640–641;<br />

teliospore surface 639, 640<br />

Ustilago avenae 637; basidium 489;<br />

teliospore germination 636<br />

Ustilago filiformis 639; teliospore<br />

germination 641, 642<br />

Ustilago hordei 639, 643; sheath 642;<br />

systemic growth 643<br />

Ustilago maydis 643–647, Pl. 12;<br />

cy<strong>to</strong>skele<strong>to</strong>n 646; dikaryon<br />

establishment 644, 645, 646; life<br />

cycle 638; mating system 605,<br />

644; mating type fac<strong>to</strong>rs 508;<br />

killer <strong>to</strong>xins 646–647; pheromones<br />

644<br />

Ustilago nuda 643; teliospore<br />

germination 641<br />

Ustilago segetum 639<br />

Ustilago tritici 642, 643<br />

vaccines against fungi 291<br />

vacuoles 4, 5, 11, 12–13, 66, 76,<br />

273, 585; in basidium<br />

development 488, 489–490; in<br />

zoospore cleaveage 87<br />

valley fever 293<br />

VAM (vesicular-arbuscular<br />

mycorrhiza) 218–221; physiology<br />

221–222<br />

Varicosporium elodeae 688, 689<br />

variegatic acid 557, 585<br />

Vascellum 581<br />

vegetative incompatibility 53, 227,<br />

365; in Basidiomycota 510, 511; in<br />

Daldinia concentrica 334; in Physarum<br />

polycephalum 48, 53; inPodospora<br />

anserina 325–326; in Trametes<br />

versicolor 511, 562<br />

Venturia 478, 481<br />

Venturia inaequalis 478, 479; control of<br />

ascospore inoculum 478–479;<br />

fungicides 479, 632<br />

veratryl alcohol 530, 531, 532<br />

vertical resistance; see host resistance<br />

vertical transmission of inoculum<br />

357<br />

Verticillium balanoides, nema<strong>to</strong>de<br />

infection process 684<br />

Verticillium lecanii 412<br />

vesicle, conidiophore 298, 302;<br />

sporangium 95, 98; sub-s<strong>to</strong>matal<br />

616; VAM 218, 220<br />

viability of propagules; see survival<br />

vic<strong>to</strong>rin C 476<br />

vines; see grape-vines<br />

viridin 339<br />

viruses, fungi as vec<strong>to</strong>rs 62, 147–148;<br />

see mycoviruses<br />

vitamin C 434<br />

Volucrispora 685<br />

volva 520<br />

Volvariella 545; life cycle 545<br />

Volvariella bombycina 545; cystidia 525;<br />

fruiting in the labora<strong>to</strong>ry 545;<br />

secotioid form 578<br />

Volvariella speciosa 543<br />

Volvariella surrecta 543, 546<br />

Volvariella volvacea 545–546;<br />

cultivation 525<br />

vomi<strong>to</strong>xin 348<br />

vulpinic acid 452, 453<br />

watercress, crook-root disease 61, 63<br />

water-moulds 79<br />

weathering of rocks 447<br />

weight-lifting by Coprinus basidiocarps<br />

538<br />

wheat, black stem rust 620; brown<br />

leaf rust 627; common bunt 647;<br />

dwarf bunt 649; eyespot 439; glume<br />

blotch 462; Karnal bunt 649;<br />

leaf blotch 462, 481, 483; leaf<br />

stripe-smut 649; loose smut 642,<br />

643; stripe rust 627; tan-spot 478<br />

whiplash flagellum; see flagella<br />

white-rot (of wood) 333, 529, 530–532,<br />

541, 560, 573, 601, Pl. 10<br />

wilts 345, 346, 366, 369<br />

wind dispersal 399<br />

wine production 275–276<br />

witches’ brooms 251<br />

wood 528; degradation 527–532<br />

Woronin bodies 227, 228, 261, 287<br />

Xanthophyllomyces dendrorhous<br />

665–666, 667; astaxanthin<br />

production 665; basidium 667;<br />

life cycle 666<br />

Xanthoria 455<br />

Xanthoria parietina 455, 456, Pl. 8; mite<br />

dispersal 448<br />

xenospore 22<br />

xerocomic acid 585<br />

Xerocomus 557<br />

xerophilic fungi 286, 298, 305<br />

Xylaria 335<br />

Xylaria carpophila 333, 335<br />

Xylaria hypoxylon 231, 334, 335, 336<br />

Xylaria longipes 335, Pl. 5; apical<br />

apparatus 241, 242<br />

Xylaria polymorpha 335<br />

Xylariales 316, 332–336


INDEX<br />

841<br />

yeasts 3; see Archiascomycetes, black<br />

yeasts, Hemiascomycetes, Mucor<br />

rouxii, Tremellales<br />

zearalenone 348<br />

Zoopagales 200–202, 675<br />

Zoophthora 203<br />

zoospore 23, 24, 57; auxiliary and<br />

principal 76, 81, 83; Chytridiomycota<br />

129–130, 131, 146, 153, 154,<br />

155, 163; cyst 84, 87, 94; Hyphochytrium<br />

70; Labyrinthulomycota 73;<br />

Oomycota 76, 77, 95; in Phy<strong>to</strong>phthora<br />

107–108; in Pythium 100; as vec<strong>to</strong>rs<br />

for viruses; see viruses, fungi as<br />

vec<strong>to</strong>rs<br />

Zostera; see eelgrass (Zostera) wasting<br />

Zygomycota 165–215, 216–225, Pl. 3;<br />

phylogeny 166<br />

zygophore 174<br />

Zygorhizidium affluens 142<br />

Zygorhizidium plank<strong>to</strong>nicum 132, 133,<br />

142<br />

Zygorhynchus 182<br />

Zygorhynchus heterogamus 172, 182<br />

Zygorhynchus moelleri 173, 175, 182,<br />

183<br />

Zygorhynchus psychrophilus 182<br />

zygosporangium 173<br />

zygospore 28, 29, 165, 174, 177, 183,<br />

186, 200, 209; formation 176–177;<br />

investment 177; germination<br />

178–179

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!