04.07.2015 Views

The genus Cladosporium and similar dematiaceous ... - CBS - KNAW

The genus Cladosporium and similar dematiaceous ... - CBS - KNAW

The genus Cladosporium and similar dematiaceous ... - CBS - KNAW

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.

Studies in Mycology 58 (2007)<br />

<strong>The</strong> <strong>genus</strong> <strong>Cladosporium</strong> <strong>and</strong> <strong>similar</strong><br />

<strong>dematiaceous</strong> hyphomycetes<br />

Pedro W. Crous, Uwe Braun, Konstanze Schubert<br />

<strong>and</strong> Johannes Z. Groenewald<br />

Centraalbureau voor Schimmelcultures,<br />

Utrecht, <strong>The</strong> Netherl<strong>and</strong>s<br />

An institute of the Royal Netherl<strong>and</strong>s Academy of Arts <strong>and</strong> Sciences


<strong>The</strong> <strong>genus</strong> <strong>Cladosporium</strong> <strong>and</strong> <strong>similar</strong> <strong>dematiaceous</strong> hyphomycetes<br />

Studies in Mycology 58, 2007


Studies in Mycology<br />

<strong>The</strong> Studies in Mycology is an international journal which publishes systematic monographs of filamentous fungi <strong>and</strong> yeasts, <strong>and</strong> in rare<br />

occasions the proceedings of special meetings related to all fields of mycology, biotechnology, ecology, molecular biology, pathology <strong>and</strong><br />

systematics. For instructions for authors see www.cbs.knaw.nl.<br />

Executive Editor<br />

Prof. dr Robert A. Samson, <strong>CBS</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD Utrecht, <strong>The</strong> Netherl<strong>and</strong>s.<br />

E-mail: r.samson@cbs.knaw.nl<br />

Layout Editor<br />

Manon van den Hoeven-Verweij, <strong>CBS</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD Utrecht, <strong>The</strong> Netherl<strong>and</strong>s.<br />

E-mail: m.verweij@cbs.knaw.nl<br />

Scientific Editors<br />

Prof. dr Uwe Braun, Martin-Luther-Universität, Institut für Biologie, Geobotanik und Botanischer Garten, Neuwerk 21,<br />

D-06099 Halle, Germany.<br />

E-mail: uwe.braun@botanik.uni-halle.de<br />

Prof. dr Pedro W. Crous, <strong>CBS</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD Utrecht, <strong>The</strong> Netherl<strong>and</strong>s.<br />

E-mail: p.crous@cbs.knaw.nl<br />

Prof. dr David M. Geiser, Department of Plant Pathology, 121 Buckhout Laboratory, Pennsylvania State University, University Park, PA, U.S.A. 16802.<br />

E-mail: dgeiser@psu.edu<br />

Dr Lorelei L. Norvell, Pacific Northwest Mycology Service, 6720 NW Skyline Blvd, Portl<strong>and</strong>, OR, U.S.A. 97229-1309.<br />

E-mail: llnorvell@pnw-ms.com<br />

Dr Erast Parmasto, Institute of Zoology & Botany, 181 Riia Street, Tartu, Estonia EE-51014.<br />

E-mail: e.parmasto@zbi.ee<br />

Prof. dr Alan J.L. Philips, Faculdade de Ciências e Tecnologia, Universidade Nova de Lisboa, Quinta de Torre, 2829-516 Caparica, Portugal.<br />

E-mail: alp@mail.fct.unl.pt<br />

Dr Amy Y. Rossman, Rm 304, Bldg 011A, Systematic Botany & Mycology Laboratory, Beltsville, MD, U.S.A. 20705.<br />

E-mail: amy@nt.ars-grin.gov<br />

Dr Keith A. Seifert, Research Scientist / Biodiversity (Mycology <strong>and</strong> Botany), Agriculture & Agri-Food Canada, KW Neatby Bldg, 960 Carling Avenue,<br />

Ottawa, ON, Canada K1A OC6.<br />

E-mail: seifertk@agr.gc.ca<br />

Prof. dr Jeffrey K. Stone, Department of Botany & Plant Pathology, Cordley 2082, Oregon State University, Corvallis, OR, U.S.A. 97331-2902.<br />

E-mail: stonej@bcc.orst.edu<br />

Dr Richard C. Summerbell, 27 Hillcrest Park, Toronto, Ont. M4X 1E8, Canada.<br />

E-mail: summerbell@aol.com<br />

Copyright 2007 Centraalbureau voor Schimmelcultures, Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD Utrecht, <strong>The</strong> Netherl<strong>and</strong>s.<br />

You are free to share — to copy, distribute <strong>and</strong> transmit the work, under the following conditions:<br />

Attribution:<br />

You must attribute the work in the manner specified by the author or licensor (but not in any way that suggests that they endorse you<br />

or your use of the work).<br />

Non-commercial: You may not use this work for commercial purposes.<br />

No derivative works: You may not alter, transform, or build upon this work.<br />

For any reuse or distribution, you must make clear to others the license terms of this work, which can be found at http://creativecommons.org/licenses/bync-nd/3.0/legalcode.<br />

Any of the above conditions can be waived if you get permission from the copyright holder. Nothing in this license impairs or restricts<br />

the author’s moral rights.<br />

Publication date: 31 August 2007<br />

Published <strong>and</strong> distributed by <strong>CBS</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD Utrecht, <strong>The</strong> Netherl<strong>and</strong>s. Internet: www.cbs.knaw.nl.<br />

E-mail: info@cbs.knaw.nl.<br />

ISBN/EAN : 978-90-70351-61-0<br />

Online ISSN : 1872-9797<br />

Print ISSN : 0166-0616<br />

Cover: Top: page 137, Fig. 30C <strong>Cladosporium</strong> pseudiridis (<strong>CBS</strong> 116463), conidiophores <strong>and</strong> conidia; page 40, Fig. 4D Toxicocladosporium irritans (type<br />

material), macroconidiophores; page 70, Fig. 11C Ramichloridium musae (<strong>CBS</strong> 365.36), sympodially proliferating conidiogenous cells, resulting in a long<br />

conidium-bearing rachis; Bottom: page 19, Fig. 8F Penidiella columbiana (type material), conidiophores with chains of disarticulating conidia; page 120, Fig.<br />

12C <strong>Cladosporium</strong> bruhnei (CPC 12211), conidial chains; page 126, Fig. 18L Davidiella tassiana (CPC 12181), asci in culture.


<strong>The</strong> <strong>genus</strong> <strong>Cladosporium</strong> <strong>and</strong> <strong>similar</strong><br />

<strong>dematiaceous</strong> hyphomycetes<br />

edited by<br />

Pedro W. Crous<br />

<strong>CBS</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD Utrecht, <strong>The</strong> Netherl<strong>and</strong>s<br />

Uwe Braun<br />

Martin-Luther-Universität Institut für Biologie, Geobotanik und Botanischer Garten, Neuwerk 21, D-06099 Halle (Saale), Germany<br />

Konstanze Schubert<br />

Botanische Staatssammlung München, Menzinger Strasse 67, D-80638 München, Germany<br />

<strong>and</strong><br />

Johannes Z. Groenewald<br />

<strong>CBS</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD Utrecht, <strong>The</strong> Netherl<strong>and</strong>s<br />

Centraalbureau voor Schimmelcultures,<br />

Utrecht, <strong>The</strong> Netherl<strong>and</strong>s<br />

An institute of the Royal Netherl<strong>and</strong>s Academy of Arts <strong>and</strong> Sciences


CONTENTS<br />

P.W. Crous, U. Braun <strong>and</strong> J.Z. Groenewald: Mycosphaerella is polyphyletic ...................................................................................................1<br />

P.W. Crous, U. Braun, K. Schubert <strong>and</strong> J.Z. Groenewald: Delimiting <strong>Cladosporium</strong> from morphologically <strong>similar</strong> genera ............................33<br />

M. Arzanlou, J.Z. Groenewald, W. Gams, U. Braun, H.-D Shin <strong>and</strong> P.W. Crous: Phylogenetic <strong>and</strong> morphotaxonomic revision of Ramichloridium<br />

<strong>and</strong> allied genera............................................................................................................................................................................................. 57<br />

K. Schubert, U. Braun, J.Z. Groenewald <strong>and</strong> P.W. Crous: <strong>Cladosporium</strong> leaf-blotch <strong>and</strong> stem rot of Paeonia spp. caused by Dichocladosporium<br />

chlorocephalum gen. nov.................................................................................................................................................................................95<br />

K. Schubert, J.Z. Groenewald, U. Braun, J. Dijksterhuis, M. Starink, C.F. Hill, P. Zalar, G.S. de Hoog <strong>and</strong> P.W. Crous: Biodiversity in<br />

the <strong>Cladosporium</strong> herbarum complex (Davidiellaceae, Capnodiales), with st<strong>and</strong>ardisation of methods for <strong>Cladosporium</strong> taxonomy<br />

<strong>and</strong> diagnostics............................................................................................................................................................................................. 105<br />

P. Zalar, G.S. de Hoog, H.-J. Schroers, P.W. Crous, J.Z. Groenewald <strong>and</strong> N. Gunde-Cimerman: Phylogeny <strong>and</strong> ecology of the<br />

ubiquitous saprobe <strong>Cladosporium</strong> sphaerospermum, with descriptions of seven new species from hypersaline environments<br />

...................................................................................................................................................................................................................... 157<br />

P.W. Crous, K. Schubert, U. Braun, G.S. de Hoog, A.D. Hocking, H.-D. Shin <strong>and</strong> J.Z. Groenewald: Opportunistic, humanpathogenic<br />

species in the Herpotrichiellaceae are phenotypically <strong>similar</strong> to saprobic or phytopathogenic species in the<br />

Venturiaceae ..............................................................................................................................................................................................185<br />

G.S. de Hoog, A.S. Nishikaku, G. Fern<strong>and</strong>ez Zeppenfeldt, C. Padín-González, E. Burger, H. Badali <strong>and</strong> A.H.G. Gerrits van den Ende:<br />

Molecular analysis <strong>and</strong> pathogenicity of the Cladophialophora carrionii complex, with the description of a novel species.......................... 219<br />

K.A. Seifert, S.J. Hughes, H. Boulay <strong>and</strong> G. Louis-Seize: Taxonomy, nomenclature <strong>and</strong> phylogeny of three cladosporiumlike<br />

hyphomycetes, Sorocybe resinae, Seifertia azalea <strong>and</strong> the Hormoconis anamorph of Amorphotheca resinae<br />

...................................................................................................................................................................................................................... 235


PREFACE<br />

DEDICATION<br />

This volume of the Studies in Mycology is dedicated to the memory<br />

of Gerardus Albertus de Vries (1919–2005), who spent his scientific<br />

career as mycologist at the <strong>CBS</strong>, where he was appointed as<br />

medical mycologist.<br />

of loyal friends <strong>and</strong> fellow mycologists dating back to the “Baarnera”<br />

of <strong>CBS</strong>. To the very end the <strong>CBS</strong> received a yearly Christmas<br />

card, which was always a water colour painting depicting some<br />

fascinating birds that he happened to be studying at the time. <strong>The</strong><br />

<strong>Cladosporium</strong> notebooks, annotations <strong>and</strong> live cultures are still at<br />

the <strong>CBS</strong>. It is thus with great joy that we dedicate this volume to<br />

Gerard de Vries, <strong>and</strong> build on his <strong>Cladosporium</strong> legacy, most of<br />

which is still sporulating, <strong>and</strong> will be available for scientific debate<br />

for generations to come.<br />

<strong>The</strong> <strong>genus</strong> <strong>Cladosporium</strong> <strong>and</strong> <strong>similar</strong> <strong>dematiaceous</strong><br />

hyphomycetes<br />

INTRODUCTION<br />

Fig. 1. Gerard de Vries studying <strong>Cladosporium</strong> spp. on different cultural growth<br />

media.<br />

On the 10 th of July 1952, Gerard de Vries graduated from the<br />

University of Utrecht, where he completed his Ph.D. on the topic<br />

“Contribution to the knowledge of the <strong>genus</strong> <strong>Cladosporium</strong> Link<br />

ex Fr.” under the guidance of the then director of the <strong>CBS</strong>, Prof.<br />

dr Johanna Westerdijk. At this time, his thesis (de Vries 1952)<br />

represented a benchmark <strong>and</strong> synthesis of our knowledge <strong>and</strong><br />

underst<strong>and</strong>ing of <strong>Cladosporium</strong> spp. studied in culture (Fig. 1).<br />

Gerard de Vries learned the basics of mycology by working during<br />

the 1930’s with mushroom taxonomy under the guidance of<br />

Abraham van Luyk. Via van Luyk he was also introduced to the<br />

Dutch Mycological Society, who organised excursions, also around<br />

Baarn, which is where the de Vries family lived. For the rest of his<br />

career, Gerard would retain his love for studying <strong>and</strong> collecting<br />

mushrooms. In 1948 he was employed at <strong>CBS</strong> under Johanna<br />

Westerdijk, <strong>and</strong> given the task of establishing <strong>and</strong> heading a<br />

new division called Medical Mycology. This he did, right up to his<br />

retirement in 1984. For several years de Vries played an important<br />

role in this discipline, <strong>and</strong> attended numerous medical congresses<br />

<strong>and</strong> workshops, <strong>and</strong> also published extensively on the topic. De<br />

Vries loved the outdoors <strong>and</strong> traveling, <strong>and</strong> combined this with his<br />

other passion, which was ornithology (van der Aa 2005). In 1952<br />

he graduated from the University of Utrecht, producing a revision<br />

of major species in the <strong>genus</strong> <strong>Cladosporium</strong> of importance to the<br />

medical, industrial <strong>and</strong> plant pathology disciplines. His book soon<br />

became highly popular, <strong>and</strong> with a strong dem<strong>and</strong> for additional<br />

copies, resulting in it being reprinted by J. Cramer. His contribution<br />

to <strong>Cladosporium</strong> was also recently acknowledged with the<br />

introduction of the cladosporium-like heat-tolerant <strong>genus</strong>, Devriesia<br />

by Seifert et al. (2004). Additional books published by de Vries dealt<br />

with mushrooms for amateur mycologists (1955) <strong>and</strong> a treatment of<br />

Hypogaea <strong>and</strong> truffels published in 1971 as part 3 of the “Fungi of<br />

the Netherl<strong>and</strong>s” (van der Aa 2005).<br />

Gerard de Vries was a well tempered, softly spoken man,<br />

who avoided conflicts, except when it dealt with scientific issues.<br />

He never married, <strong>and</strong> died a bachelor, surrounded by a circle<br />

Species of <strong>Cladosporium</strong> are common <strong>and</strong> widespread, <strong>and</strong> interact<br />

with humans in every phase of life, from growing behind your bed<br />

or bedroom cupboard <strong>and</strong> producing allergens, or growing on the<br />

bathroom ceiling, to the fruit decay happening in the fruit basket in<br />

the kitchen, to colonising the debris lying outside your house, <strong>and</strong><br />

even the plant diseases observed on some of the shrubs, trees, or<br />

flowers cultivated in your garden. However, scientists generally shy<br />

away from trying to identify these <strong>similar</strong> looking organisms, <strong>and</strong><br />

therefore the main aims of this volume were to:<br />

1) Establish st<strong>and</strong>ardised conditions <strong>and</strong> protocols for studying<br />

cladosporioid species <strong>and</strong> their teleomorphs in culture;<br />

2) Determine how they can morphologically be distinguished in<br />

culture, <strong>and</strong> highlight important diagnostic features;<br />

3) Circumscribe the <strong>genus</strong>, <strong>and</strong> delineate it from morphologically<br />

<strong>similar</strong> <strong>dematiaceous</strong> hyphomycetes;<br />

4) Determine which DNA gene loci are informative to accurately<br />

distinguish species of <strong>Cladosporium</strong>, <strong>and</strong> initiate a database of<br />

<strong>Cladosporium</strong> sequences that can in future be used to set up an<br />

online polyphasic identification key.<br />

Although <strong>Cladosporium</strong> is one of the largest <strong>and</strong> most<br />

heterogeneous genera of hyphomycetes, currently comprising<br />

more than 772 names (Dugan et al. 2004), only a mere fraction of<br />

these species are known from culture, <strong>and</strong> thus the real number<br />

of taxa that exist remains unknown. Species of <strong>Cladosporium</strong> are<br />

commonly encountered on plant <strong>and</strong> other kinds of debris, frequently<br />

colonising lesions of plant pathogenic fungi, <strong>and</strong> are also isolated<br />

from air, soil, food, paint, textiles <strong>and</strong> other organic matters (Ellis<br />

1971, 1976; Schubert 2005a), they are also common endophytes<br />

(Brown et al. 1998, El-Morsy 2000) as well as phylloplane fungi<br />

(Islam & Hasin 2000, de Jager et al. 2001, Inacio et al. 2002,<br />

Stohr & Dighton 2004, Levetin & Dorsey 2006). Some species of<br />

<strong>Cladosporium</strong> have a medical relevance in clinical laboratories,<br />

<strong>and</strong> also cause allergic lung mycoses (de Hoog et al. 2000). In<br />

spite of its obvious importance, species of <strong>Cladosporium</strong> are still<br />

poorly understood.<br />

Taxonomy of the anamorph<br />

<strong>The</strong> first binominal introduced for this group of fungi was that of<br />

Dematium herbarum Pers. (Persoon 1794) (Fig. 2). <strong>Cladosporium</strong><br />

herbarum (Pers.) Link was subsequently selected to serve as


lectotype for the <strong>genus</strong> by Clements & Shear (1931), a proposal<br />

which was accepted by de Vries (1952), Hughes (1958), <strong>and</strong> others<br />

(Prasil & de Hoog 1988). In subsequent years the number of taxa<br />

described in the <strong>genus</strong> grew rapidly, though the generic concept<br />

was rather vague. As a consequence, numerous morphologically<br />

<strong>similar</strong> <strong>dematiaceous</strong> hyphomycetes with catenulate conidia were<br />

incorrectly assigned to <strong>Cladosporium</strong>, making this one of the largest<br />

genera of hyphomycetes.<br />

In an attempt to circumscribe some of the more well-known taxa,<br />

de Vries (1952) published a revision of nine <strong>Cladosporium</strong> species in<br />

vivo <strong>and</strong> in vitro, <strong>and</strong> 13 additional taxa in an appendix. Ellis (1971,<br />

1976) described <strong>and</strong> illustrated 43 species, while Morgan-Jones<br />

<strong>and</strong> McKemy dealt with selected species in the series “Studies in<br />

the <strong>genus</strong> <strong>Cladosporium</strong> s. lat.” (Morgan-Jones & McKemy 1990,<br />

McKemy & Morgan-Jones 1990, 1991a–c). Other significant works<br />

that also treated <strong>Cladosporium</strong> species include Ho et al. (1999),<br />

<strong>and</strong> Zhang et al. (2003), though these authors still followed a<br />

wider generic concept. David (1997) followed the taxonomy of de<br />

Vries (1952), who first considered Heterosporium as a synonym<br />

of <strong>Cladosporium</strong>, <strong>and</strong> introduced the combination <strong>Cladosporium</strong><br />

subgen. Heterosporium. Subsequent to this publication, further<br />

monographic studies on the <strong>genus</strong> <strong>Cladosporium</strong> s. lat. were<br />

initiated by Braun <strong>and</strong> co-workers (Braun et al. 2003, 2006, Dugan<br />

et al. 2004, Schubert & Braun 2004, 2005a, b, 2006, 2007, Schubert<br />

2005a, b, Heuchert et al. 2005).<br />

Treatments of human pathogenic <strong>Cladosporium</strong> species<br />

(Masclaux et al. 1995, Untereiner 1997, Gerrits van den Ende & de<br />

Hoog 1999, Untereiner & Naveau 1999, Untereiner et al. 1999; de<br />

Hoog et al. 2000), concluded that they represent species belonging<br />

to the Herpotrichiellaceae (Capronia Sacc./Cladophialophora<br />

Borelli). Saprobic species, which appear morphologically <strong>similar</strong>,<br />

were found to belong to the Venturiaceae (Caproventuria U. Braun/<br />

Pseudocladosporium U. Braun; Braun et al. 2003, Schubert et al.<br />

2003, Beck et al. 2005) (see Crous et al. 2007 – this volume). Further<br />

genera that were separated from <strong>Cladosporium</strong> include Sorocybe<br />

resinae (Fr.) Fr. [≡ <strong>Cladosporium</strong> resinae (Lindau) G.A. de Vries,<br />

teleomorph: Amorphotheca resinae Parbery; Partridge & Morgan-<br />

Jones 2002] (see Seifert et al. 2007 – this volume), Devriesia Seifert<br />

& N.L. Nickerson, erected for heat tolerant species (Seifert et al.<br />

2004), Cladoriella Crous, erected for saprobic species (Crous et<br />

al. 2006b), Metulocladosporiella Crous, Schroers, Groenewald, U.<br />

Braun & K. Schub., erected for the causal agent of banana speckle<br />

disease (Crous et al. 2006a), Digitopodium U. Braun, Heuchert &<br />

K. Schub. <strong>and</strong> Parapericoniella U. Braun, Heuchert & K. Schub.,<br />

Fig. 2. Type specimen of Dematium herbarum Pers. (1794), preserved in the<br />

National Herbarium of the Netherl<strong>and</strong>s in Leiden.<br />

representing two genera of hyperparasitic hyphomycetes (Heuchert<br />

et al. 2005).<br />

Taxonomy of the teleomorph<br />

Teleomorphs of <strong>Cladosporium</strong> have traditionally been described in<br />

Mycosphaerella Johanson. <strong>The</strong> first indication that this may not be<br />

the case was the rDNA ITS sequence data presented by Crous et<br />

al. (2001), which revealed cladosporium-like taxa to cluster basal<br />

to Mycosphaerella s. str. This finding was further strengthened by<br />

adding 18S rDNA data, which clearly distinguished the <strong>Cladosporium</strong><br />

clade from Mycosphaerella (Braun et al. 2003). <strong>The</strong>se results<br />

lead to the erection of the <strong>genus</strong> Davidiella Crous & U. Braun for<br />

teleomorphs of <strong>Cladosporium</strong>, though it was largely established<br />

based on its unique anamorphs, rather than distinct teleomorph<br />

features. In a revision of the <strong>genus</strong> Mycosphaerella, Aptroot (2006)<br />

provided the first clear morphological characteristics to distinguish<br />

Davidiella from Mycosphaerella, referring to their sole-shaped<br />

ascospores, <strong>and</strong> angular lumina that are to be seen in Davidiella<br />

ascospores. Further phylogenetic evidence for the distinction was<br />

found by Schoch et al. (2006), which led to the erection of the<br />

family Davidiellaceae (Capnodiales). Detailed cultural studies of<br />

Davidiella teleomorphs, however, were still lacking (see Schubert<br />

et al. 2007 – this volume).<br />

What is <strong>Cladosporium</strong>?<br />

David (1997) provided the first modern concept of <strong>Cladosporium</strong> by<br />

conducting comprehensive scanning electron microscopic (SEM)<br />

examinations of the scar <strong>and</strong> hilum structure in <strong>Cladosporium</strong> <strong>and</strong><br />

Heterosporium, thereby confirming the observations of Roquebert<br />

(1981). He introduced the term “coronate” for the <strong>Cladosporium</strong><br />

scar type, which is characterised by having a central convex part<br />

(dome), surrounded by a raised periclinal rim (Fig. 3), <strong>and</strong> proved<br />

that these anamorphs are linked to teleomorphs now placed in<br />

Davidiella (see David 1997, fig. 12).<br />

This new concept of <strong>Cladosporium</strong> s. str. <strong>and</strong> Davidiella<br />

(David 1997, Braun et al. 2003, Aptroot 2006), supported by<br />

morphological <strong>and</strong> molecular data, rendered it possible to initiate<br />

a comprehensive revision of the <strong>genus</strong>. <strong>The</strong> first step was the<br />

preparation of a general, annotated check-list of <strong>Cladosporium</strong><br />

names (Dugan et al. 2004), followed by revisions of fungicolous<br />

(Heuchert et al. 2005) <strong>and</strong> foliicolous species of <strong>Cladosporium</strong> s.<br />

lat. (Schubert 2005b, Braun et al. 2006, Schubert & Braun 2004,<br />

2005a, b, 2006, 2007). <strong>The</strong> present study is the first to integrate<br />

these concepts on cladosporioid species in culture, in an attempt to<br />

further elucidate species of <strong>Cladosporium</strong>, <strong>and</strong> delineate the <strong>genus</strong><br />

from other, morphologically <strong>similar</strong> <strong>dematiaceous</strong> genera that have<br />

traditionally been confused with <strong>Cladosporium</strong> s. str.<br />

How natural should anamorph genera be?<br />

Article 59 of the International Code of Botanical Nomenclature<br />

was introduced to enable mycologists to name the asexual states<br />

of fungi that they encountered, <strong>and</strong> for which no teleomorph<br />

association was known. It was <strong>and</strong> remains a completely artificial<br />

system, complicated further by the evolution of the same anamorph<br />

morphology in different families, <strong>and</strong> even orders. In 1995 Gams<br />

discussed “How natural should anamorph genera be”, concluding<br />

that paraphyletic genera should be an acceptable option, <strong>and</strong> that<br />

anamorphs cannot reflect natural relationships. In a special volume<br />

dedicated at integrating molecular data <strong>and</strong> morphology, Seifert et<br />

al. (2000) proposed using anamorph names as adjectives, e.g.,<br />

acremonium-like, when they clustered in different clades, or were<br />

linked to different teleomorphs than the type species of the <strong>genus</strong>


Fig. 3. Coronate scar structure of <strong>Cladosporium</strong> herbaroides, visible by means of<br />

Scanning Electron Microscopy. Scale bar = 5 µm (Photo: Jan Dijksterhuis).<br />

Acremonium. In a phylogenetic study of the Herpotrichiellaceae,<br />

Haase et al. (1999) proposed to accept anamorphs as poly- <strong>and</strong><br />

paraphyletic within the order Chaetothyriales, as their taxonomy<br />

was unsupported by phylogeny, <strong>and</strong> Cook et al. (1997) as well as<br />

Braun et al. (2002) followed this methodology in naming anamorphs<br />

of the Erysiphaceae (Erysiphales). After much debate, we have<br />

chosen to use the same approach in this volume, <strong>and</strong> will refrain<br />

from introducing different anamorph genera for the same phenotype<br />

clustering within different clades of the same order. It is hoped that<br />

this approach will stop the unnecessary proliferation of names, until<br />

we can move to a single nomenclature for ascomycetous fungi.<br />

Anamorphs are just form taxa, established for the sole purpose<br />

of enabling mycologists to name asexual states that occur in<br />

the absence of their teleomorphs. Anamorph genera are simply<br />

phenotypic concepts that lack phylogenetic relevance within the<br />

order.<br />

REFERENCES<br />

Aa HA van der (2005). In memoriam Gerard de Vries (1919–2005). Coolia 48(4):<br />

173–175.<br />

Aptroot A (2006). Mycosphaerella <strong>and</strong> its anamorphs: 2. Conspectus of<br />

Mycosphaerella. <strong>CBS</strong> Biodiversity Series 5: 1–231.<br />

Beck A, Ritschel A, Schubert K, Braun U, Triebel D (2005). Phylogenetic relationship<br />

of the anamorphic <strong>genus</strong> Fusicladium s. lat. as inferred by ITS data. Mycological<br />

Progress 4: 111–116.<br />

Braun U, Cook RTA, Inman AJ, Shin HD (2002). <strong>The</strong> taxonomy of the powdery<br />

mildew fungi. In: <strong>The</strong> powdery mildews, a comprehensive treatise (Bélanger<br />

RR, Bushnell WR, Dik AJ, Carver TLW, eds). APS Press, St. Paul, U.S.A.:<br />

13–55.<br />

Braun U, Crous PW, Dugan FM, Groenewald JZ, Hoog GS de (2003). Phylogeny<br />

<strong>and</strong> taxonomy of cladosporium-like hyphomycetes, including Davidiella gen.<br />

nov., the teleomorph of <strong>Cladosporium</strong> s. str. Mycological Progress 2: 3–18.<br />

Braun U, Hill CF, Schubert K (2006). New species <strong>and</strong> new records of biotrophic<br />

micromycetes from Australia, Fiji, New Zeal<strong>and</strong> <strong>and</strong> Thail<strong>and</strong>. Fungal Diversity<br />

22: 13–35.<br />

Brown KB, Hyde KD, Guest DI (1998). Preliminary studies on endophytic fungal<br />

communities of Musa acuminata species complex in Hong Kong <strong>and</strong> Australia.<br />

Fungal Diversity 1: 27–51.<br />

Clements FE, Shear CL (1931). <strong>The</strong> Genera of Fungi. H.W. Wilson, New York.<br />

Cook RTA, Inman AJ, Billings C (1997). Identification <strong>and</strong> classification of powdery<br />

mildew anamorphs using light <strong>and</strong> scanning electron microscopy <strong>and</strong> host<br />

range data. Mycological Research 101: 975–1002.<br />

Crous PW, Kang JC, Braun U (2001). A phylogenetic redefinition of anamorph<br />

genera in Mycosphaerella based on ITS rDNA sequences. Mycologia 93:<br />

1081–1101.<br />

Crous PW, Schroers HJ, Groenewald JZ, Braun U, Schubert K (2006a).<br />

Metulocladosporiella gen. nov. for the causal organism of <strong>Cladosporium</strong><br />

speckle disease of banana. Mycological Research 110: 264–275.<br />

Crous PW, Schubert K, Braun U, de Hoog GS, Hocking AD, Shin H-D, Groenewald<br />

JZ (2007). Opportunistic, human-pathogenic species in the Herpotrichiellaceae<br />

are phenotypically <strong>similar</strong> to saprobic or phytopathogenic species in the<br />

Venturiaceae. Studies in Mycology 58: 185–217.<br />

Crous PW, Verkley GJM, Groenewald JZ (2006b). Eucalyptus microfungi known<br />

from culture. 1. Cladoriella <strong>and</strong> Fulvoflamma genera nova, with notes on some<br />

other poorly known taxa. Studies in Mycology 55: 53–63.<br />

David JC (1997). A contribution to the systematics of <strong>Cladosporium</strong>. Revision of the<br />

fungi previously referred to Heterosporium. Mycological Papers 172: 1–157.<br />

Dugan FM, Schubert K, Braun U (2004). Check-list of <strong>Cladosporium</strong> names.<br />

Schlechtendalia 11: 1–103.<br />

El-Morsy EM (2000). Fungi isolated from the endorhizosphere of halophytic plants<br />

from the Red Sea Coast of Egypt. Fungal Diversity 5: 43–54.<br />

Ellis MB (1971). Dematiaceous Hyphomycetes. Commonwealth Mycological<br />

Institute, Kew.<br />

Ellis MB (1976). More Dematiaceous Hyphomycetes. Commonwealth Mycological<br />

Institute, Kew.<br />

Gams W (1995). How natural should anamorph genera be? Canadian Journal of<br />

Botany 73: S747–S753.<br />

Gerrits van den Ende AHG, Hoog GS de (1999). Variability <strong>and</strong> molecular diagnostics<br />

of the neurotrophic species Cladophialophora bantiana. Studies in Mycology<br />

43: 151–162.<br />

Haase G, Sonntag L, Melzer-Krick B, Hoog GS de (1999). Phylogenetic inference<br />

by SSU-gene analysis of members of the Herpotrichiellaceae with special<br />

reference to human pathogenic species. Studies in Mycology 43: 80–97.<br />

Heuchert B, Braun U, Schubert K (2005). Morphotaxonomic revision of fungicolous<br />

<strong>Cladosporium</strong> species (hyphomycetes). Schlechtendalia 13: 1–78.<br />

Ho MH-M, Castañeda RF, Dugan FM, Jong, SC (1999). <strong>Cladosporium</strong> <strong>and</strong><br />

Cladophialophora in culture: descriptions <strong>and</strong> an exp<strong>and</strong>ed key. Mycotaxon<br />

72: 115–157.<br />

Hoog GS de, Guarro, J, Gené J, Figueras MJ (2000). Atlas of clinical fungi. 2 nd ed.<br />

Centralbureau voor Schimmelcultures, Utrecht <strong>and</strong> Universitat rovira I virgili,<br />

Reus.<br />

Hughes SJ (1958). Revisiones hyphomycetum aliquot cum appendice de nominibus<br />

rejiciendis. Canadian Journal of Botany 36: 727–836.<br />

Inacio J, Pereira P, de Cavalho M, Fonseca A, Amaral-Collaco MT, Spencer-Martins<br />

I (2002). Estimation <strong>and</strong> diversity of phylloplane mycobiota on selected plants in<br />

a Mediterranean-type ecosystem in Portugal. Microbial Ecology 44: 344–353.<br />

Islam M, Hasin F (2000). Studies on phylloplane mycoflora of Amaranthus viridis L.<br />

National Academy Science Letters, India 23: 121–123.<br />

Jager ES de, Wehner FC, Karsten L (2001). Microbial ecology of the mango<br />

phylloplane. Microbial Ecology 42: 201–207.<br />

Levetin E, Dorsey K (2006). Contribution to leaf surface fungi to the air spora.<br />

Aerobiologia 22: 3–12.<br />

Masclaux F, Guého E, Hoog GS de, Christen R (1995). Phylogenetic relationship of<br />

human-pathogenic <strong>Cladosporium</strong> (Xylohypha) species. Journal of Medical <strong>and</strong><br />

Veterinary Mycology 33: 327–338.<br />

McKemy JM, Morgan-Jones G (1990). Studies in the <strong>genus</strong> <strong>Cladosporium</strong> sensu<br />

lato II. Concerning Heterosporium gracile, the causal organism of leaf spot<br />

disease of Iris species. Mycotaxon 39: 425–440.<br />

McKemy JM, Morgan-Jones G (1991a). Studies in the <strong>genus</strong> <strong>Cladosporium</strong><br />

sensu lato III. Concerning <strong>Cladosporium</strong> chlorocephalum <strong>and</strong> its synonym<br />

<strong>Cladosporium</strong> paeoniae, the causal organism of leaf-blotch of peony.<br />

Mycotaxon 41: 135–146.<br />

McKemy JM, Morgan-Jones G (1991b). Studies in the <strong>genus</strong> <strong>Cladosporium</strong> sensu<br />

lato IV. Concerning <strong>Cladosporium</strong> oxysporum, a plurivorous predominantly<br />

saprophytic species in warm climates. Mycotaxon 41: 397–405.<br />

McKemy JM, Morgan-Jones G (1991c). Studies in the <strong>genus</strong> <strong>Cladosporium</strong> sensu<br />

lato V. Concerning the type species <strong>Cladosporium</strong> herbarum. Mycotaxon 42:<br />

307–317.<br />

Morgan-Jones G, McKemy JM (1990). Studies in the <strong>genus</strong> <strong>Cladosporium</strong> sensu<br />

lato I. Concerning <strong>Cladosporium</strong> uredinicola, occurring on telial columns of<br />

Cronartium quercuum <strong>and</strong> other rusts. Mycotaxon 39: 185–202.<br />

Partridge EC, Morgan-Jones G (2002). Notes on hyphomycetes, LXXXVIII. New<br />

genera in which to classify Alysidium resinae <strong>and</strong> Pycnostysanus azaleae, with<br />

a consideration of Sorocybe. Mycotaxon 83: 335–352.<br />

Persoon CH (1794). Neuer Versuch einer systematischen Eintheilung der<br />

Schwämme. Neues Magazin fur die Botanik 1: 63–128.<br />

Prasil K, Hoog GS de (1988). Variability in <strong>Cladosporium</strong> herbarum. Transactions of<br />

the British Mycological Society 90: 49–54.<br />

Roquebert MF (1981). Analyse des phénoménes pariétaux au cours de la<br />

conidiogenése chez quelques champignon microscopiques. Memoires du<br />

Museum d’Histoire Naturelle, Sér B, Botanique 28: 3–79.<br />

Schoch C, Shoemaker RA, Seifert KA, Hambleton S, Spatafora JW, Crous PW<br />

(2006). A multigene phylogeny of the Dothideomycetes using four nuclear loci.<br />

Mycologia 98: 1041–1052.


Schubert K (2005a). Morphotaxonomic revision of foliicolous <strong>Cladosporium</strong> species<br />

(hyphomycetes). PhD thesis, Martin-Luther-University, Halle.<br />

Schubert K (2005b). Taxonomic revision of the <strong>genus</strong> <strong>Cladosporium</strong> s. lat. 3. A<br />

revision of <strong>Cladosporium</strong> species described by J.J. Davis <strong>and</strong> H.C. Greene<br />

(WIS). Mycotaxon 92: 55–76.<br />

Schubert K, Braun U (2004). Taxonomic revision of the <strong>genus</strong> <strong>Cladosporium</strong> s. lat. 2.<br />

Morphotaxonomic examination of <strong>Cladosporium</strong> species occurring on hosts of<br />

the families Bignoniaceae <strong>and</strong> Orchidaceae. Sydowia 56: 76–97.<br />

Schubert K, Braun U (2005a). Taxonomic revision of the <strong>genus</strong> <strong>Cladosporium</strong><br />

s. lat. 1. Species reallocated to Fusicladium, Parastenella, Passalora,<br />

Pseudocercospora <strong>and</strong> Stenella. Mycological Progress 4: 101–109.<br />

Schubert K, Braun U (2005b). Taxonomic revision of the <strong>genus</strong> <strong>Cladosporium</strong> s.<br />

lat. 4. Species reallocated to Asperisporium, Dischloridium, Fusicladium,<br />

Passalora, Pseudasperisporium <strong>and</strong> Stenella. Fungal Diversity 20: 187–208.<br />

Schubert K, Braun U (2006). Taxonomic revision of the <strong>genus</strong> <strong>Cladosporium</strong> s. lat. 5.<br />

Validation <strong>and</strong> description of new species. Schlechtendalia 14: 55–83.<br />

Schubert K, Braun U (2007). Taxonomic revision of the <strong>genus</strong> <strong>Cladosporium</strong> s. lat.<br />

6. New species, reallocations to <strong>and</strong> synonyms of Cercospora, Fusicladium,<br />

Passalora, Septonema <strong>and</strong> Stenella. Nova Hedwigia 84(1–2): 189–208.<br />

Schubert K, Groenewald JZ, Braun U, Dijksterhuis J, Starink M, Hill CF, Zalar P,<br />

Hoog GS de, Crous PW (2007). Biodiversity in the <strong>Cladosporium</strong> herbarum<br />

complex (Davidiellaceae, Capnodiales), with st<strong>and</strong>ardisation of methods for<br />

<strong>Cladosporium</strong> taxonomy <strong>and</strong> diagnostics. Studies in Mycology 58: 105–156.<br />

Schubert K, Ritschel A, Braun U (2003). A monograph of Fusicladium s. lat.<br />

(hyphomycetes). Schlechtendalia 9: 1–132.<br />

Seifert KA, Gams W, Crous PW, Samuels GJ (2000). Molecules, morphology <strong>and</strong><br />

classification: Towards monophyletic genera in the Ascomycetes. Studies in<br />

Mycology 45: 1–230.<br />

Seifert KA, Hughes SJ, Boulay H, Louis-Seize G (2007). Taxonomy, nomenclature<br />

<strong>and</strong> phylogeny of three cladosporium-like hyphomycetes, Sorocybe resinae,<br />

Seifertia azalea <strong>and</strong> the Hormoconis anamorph of Amorphotheca resinae.<br />

Studies in Mycology 58: 235–245.<br />

Seifert KA, Nickerson NL, Corlett M, Jackson ED, Lois-Seize G, Davies RJ (2004).<br />

Devriesia, a new hyphomycete <strong>genus</strong> to accommodate heat-resistant,<br />

cladosporium-like fungi. Canadian Journal of Botany 82: 914–926.<br />

Stohr SN, Dighton J (2004). Effects of species diversity on establishment <strong>and</strong><br />

coexistence: A phylloplane fungal community model system. Microbial Ecology<br />

48: 431–438.<br />

Untereiner WA (1997). Taxonomy of selected members of the ascomycete <strong>genus</strong><br />

Capronia with notes on anamorph-teleomorph connection. Mycologia 89:<br />

120–131.<br />

Untereiner WA, Gerrits van den Ende AHG, Hoog GA de (1999). Nutritional<br />

physiology of species of Capronia. Studies in Mycology 43: 98–106.<br />

Untereiner WA, Naveau FA (1999). Molecular systematic of the Herpotrichiellaceae<br />

with an assessment of the phylogenetic position of Exophiala dermatitidis <strong>and</strong><br />

Phialophora americana. Mycologia 91: 67–83.<br />

Vries GA de (1952). Contribution to the Knowledge of the Genus <strong>Cladosporium</strong> Link<br />

ex Fr. Centraalbureau voor Schimmelcultures, Baarn.<br />

Vries GA de (1955). Paddestoelen (en schimmels). WJ Thieme & Cie. Zutputen.<br />

Vries GA de (1971). De Fungi van Nederl<strong>and</strong>. 3. Hypogaea. Wetenschappelijke<br />

mededelingen van de Koninklijke Nederl<strong>and</strong>se Natuurhistorische Vereniging.<br />

Hoogwoud.<br />

Zhang ZY, Liu YL, Zhang T, Li TF, Wang G, Zhang H, He YH, Peng HH (2003).<br />

<strong>Cladosporium</strong>, Fusicladium, Pyricularia. Flora Fungorum Sinicorum 14: 1–<br />

297.<br />

<strong>The</strong> Editors 1 July 2007<br />

Propositions for this volume:<br />

“Evolution gives rise to lineages, which we try to recognise as genera <strong>and</strong> species”<br />

“<strong>The</strong> most interesting fungi are those isolated by accident”


available online at www.studiesinmycology.org<br />

doi:10.3114/sim.2007.58.01<br />

Studies in Mycology 58: 1–32. 2007.<br />

Mycosphaerella is polyphyletic<br />

P.W. Crous 1* , U. Braun 2 <strong>and</strong> J.Z. Groenewald 1<br />

1<br />

<strong>CBS</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD, Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; 2 Martin-Luther-Universität, Institut für Biologie, Geobotanik und Botanischer Garten,<br />

Herbarium, Neuwerk 21, D-06099 Halle, Germany<br />

*Correspondence: Pedro W. Crous, p.crous@cbs.knaw.nl<br />

Abstract: Mycosphaerella, one of the largest genera of ascomycetes, encompasses several thous<strong>and</strong> species <strong>and</strong> has anamorphs residing in more than 30 form genera.<br />

Although previous phylogenetic studies based on the ITS rDNA locus supported the monophyly of the <strong>genus</strong>, DNA sequence data derived from the LSU gene distinguish several<br />

clades <strong>and</strong> families in what has hitherto been considered to represent the Mycosphaerellaceae. Several important leaf spotting <strong>and</strong> extremotolerant species need to be disposed<br />

to the <strong>genus</strong> Teratosphaeria, for which a new family, the Teratosphaeriaceae, is introduced. Other distinct clades represent the Schizothyriaceae, Davidiellaceae, Capnodiaceae,<br />

<strong>and</strong> the Mycosphaerellaceae. Within the two major clades, namely Teratosphaeriaceae <strong>and</strong> Mycosphaerellaceae, most anamorph genera are polyphyletic, <strong>and</strong> new anamorph<br />

concepts need to be derived to cope with dual nomenclature within the Mycosphaerella complex.<br />

Taxonomic novelties: Batcheloromyces eucalypti (Alcorn) Crous & U. Braun, comb. nov., Catenulostroma Crous & U. Braun, gen. nov., Catenulostroma abietis (Butin &<br />

Pehl) Crous & U. Braun, comb. nov., Catenulostroma chromoblastomycosum Crous & U. Braun, sp. nov., Catenulostroma elginense (Joanne E. Taylor & Crous) Crous & U.<br />

Braun, comb. nov., Catenulostroma excentricum (B. Sutton & Ganap.) Crous & U. Braun, comb. nov., Catenulostroma germanicum Crous & U. Braun, sp. nov., Catenulostroma<br />

macowanii (Sacc.) Crous & U. Braun, comb. nov., Catenulostroma microsporum (Joanne E. Taylor & Crous) Crous & U. Braun, comb. nov., Catenulostroma protearum (Crous<br />

& M.E. Palm) Crous & U. Braun, comb. nov., Penidiella Crous & U. Braun, gen. nov., Penidiella columbiana Crous & U. Braun, sp. nov., Penidiella cubensis (R.F. Castañeda) U.<br />

Braun, Crous & R.F. Castañeda, comb. nov., Penidiella nect<strong>and</strong>rae Crous, U. Braun & R.F. Castañeda, nom. nov., Penidiella rigidophora Crous, R.F. Castañeda & U. Braun, sp.<br />

nov., Penidiella strumelloidea (Milko & Dunaev) Crous & U. Braun, comb. nov., Penidiella venezuelensis Crous & U. Braun, sp. nov., Readeriella blakelyi (Crous & Summerell)<br />

Crous & U. Braun, comb. nov., Readeriella brunneotingens Crous & Summerell, sp. nov., Readeriella considenianae (Crous & Summerell) Crous & U. Braun, comb. nov.,<br />

Readeriella destructans (M.J. Wingf. & Crous) Crous & U. Braun, comb. nov., Readeriella dimorpha (Crous & Carnegie) Crous & U. Braun, comb. nov., Readeriella epicoccoides<br />

(Cooke & Massee) Crous & U. Braun, comb. nov., Readeriella gauchensis (M.-N. Cortinas, Crous & M.J. Wingf.) Crous & U. Braun, comb. nov., Readeriella molleriana (Crous<br />

& M.J. Wingf.) Crous & U. Braun, comb. nov., Readeriella nubilosa (Ganap. & Corbin) Crous & U. Braun, comb. nov., Readeriella pulcherrima (Gadgil & M. Dick) Crous &<br />

U. Braun, comb. nov., Readeriella stellenboschiana (Crous) Crous & U. Braun, comb. nov., Readeriella toledana (Crous & Bills) Crous & U. Braun, comb. nov., Readeriella<br />

zuluensis (M.J. Wingf., Crous & T.A. Cout.) Crous & U. Braun, comb. nov., Teratosphaeria africana (Crous & M.J. Wingf.) Crous & U. Braun, comb. nov., Teratosphaeria alistairii<br />

(Crous) Crous & U. Braun, comb. nov., Teratosphaeria associata (Crous & Carnegie) Crous & U. Braun, comb. nov., Teratosphaeria bellula (Crous & M.J. Wingf.) Crous & U.<br />

Braun, comb. nov., Teratosphaeria cryptica (Cooke) Crous & U. Braun, comb. nov., Teratosphaeria dentritica (Crous & Summerell) Crous & U. Braun, comb. nov., Teratosphaeria<br />

excentrica (Crous & Carnegie) Crous & U. Braun, comb. nov., Teratosphaeria fimbriata (Crous & Summerell) Crous & U. Braun, comb. nov., Teratosphaeria flexuosa (Crous<br />

& M.J. Wingf.) Crous & U. Braun, comb. nov., Teratosphaeria gamsii (Crous) Crous & U. Braun, comb. nov., Teratosphaeria jonkershoekensis (P.S. van Wyk, Marasas &<br />

Knox-Dav.) Crous & U. Braun, comb. nov., Teratosphaeria maxii (Crous) Crous & U. Braun, comb. nov., Teratosphaeria mexicana (Crous) Crous & U. Braun, comb. nov.,<br />

Teratosphaeria molleriana (Thüm.) Crous & U. Braun, comb. nov., Teratosphaeria nubilosa (Cooke) Crous & U. Braun, comb. nov., Teratosphaeria ohnowa (Crous & M.J. Wingf.)<br />

Crous & U. Braun, comb. nov., Teratosphaeria parkiiaffinis (Crous & M.J. Wingf.) Crous & U. Braun, comb. nov., Teratosphaeria parva (R.F. Park & Keane) Crous & U. Braun,<br />

comb. nov., Teratosphaeria perpendicularis (Crous & M.J. Wingf.) Crous & U. Braun, comb. nov., Teratosphaeria pluritubularis (Crous & Mansilla) Crous & U. Braun, comb.<br />

nov., Teratosphaeria pseudafricana (Crous & T.A. Cout.) Crous & U. Braun, comb. nov., Teratosphaeria pseudocryptica (Crous) Crous & U. Braun, comb. nov., Teratosphaeria<br />

pseudosuberosa (Crous & M.J. Wingf.) Crous & U. Braun, comb. nov., Teratosphaeria quasicercospora (Crous & T.A. Cout.) Crous & U. Braun, comb. nov., Teratosphaeria<br />

readeriellophora (Crous & Mansilla) Crous & U. Braun, comb. nov., Teratosphaeria secundaria (Crous & Alfenas) Crous & U. Braun, comb. nov., Teratosphaeria stramenticola<br />

(Crous & Alfenas) Crous & U. Braun, comb. nov., Teratosphaeria suberosa (Crous, F.A. Ferreira, Alfenas & M.J. Wingf.) Crous & U. Braun, comb. nov., Teratosphaeria suttonii<br />

(Crous & M.J. Wingf.) Crous & U. Braun, comb. nov., Teratosphaeria toledana (Crous & Bills) Crous & U. Braun, comb. nov., Teratosphaeriaceae Crous & U. Braun, fam. nov.<br />

Key words: Ascomycetes, Batcheloromyces, Colletogloeopsis, Readeriella, Teratosphaeria, Trimmatostroma, DNA sequence comparisons, systematics.<br />

Introduction<br />

<strong>The</strong> <strong>genus</strong> Mycosphaerella Johanson as presently circumscribed<br />

contains close to 3 000 species (Aptroot 2006), excluding its<br />

anamorphs, which represent thous<strong>and</strong>s of additional species<br />

(Crous et al. 2000, 2001, 2004a, b, 2006a, b, 2007b, Crous &<br />

Braun 2003). Crous (1998) predicted that Mycosphaerella would<br />

eventually be split according to its anamorph genera, <strong>and</strong> Crous<br />

et al. (2000) recognised six sections, as originally defined by Barr<br />

(1972). This was followed by a set of papers (Crous et al. 2001,<br />

Goodwin et al. 2001), where it was concluded, based on ITS DNA<br />

sequence data, that Mycosphaerella was monophyletic. A revision<br />

of the various coelomycete <strong>and</strong> hyphomycete anamorph concepts<br />

led Crous & Braun (2003) to propose a system whereby the asexual<br />

morphs could be allocated to various form genera affiliated with<br />

Mycosphaerella holomorphs.<br />

In a recent study that formed part of the US “Assembling<br />

the Fungal Tree of Life” project, Schoch et al. (2006) were able<br />

to show that the Mycosphaerellaceae represents a family within<br />

Capnodiales. Furthermore, some variation was also depicted within<br />

the family, which supported <strong>similar</strong> findings in other recent papers<br />

employing LSU sequence data, such as Hunter et al. (2006), <strong>and</strong><br />

Batzer et al. (2007). To further elucidate the phylogenetic variation<br />

observed within the Mycosphaerellaceae in these studies, a<br />

subset of isolates was selected for the present study, representing<br />

the various species recognised as morphologically distinct from<br />

Mycosphaerella s. str.<br />

<strong>The</strong> <strong>genus</strong> Mycosphaerella has in recent years been linked<br />

to approximately 30 anamorph genera (Crous & Braun 2003,<br />

Crous et al. 2007b). Many of these anamorph genera resulted<br />

from a reassessment of cercosporoid forms. Chupp (1954) was<br />

of the opinion that they all represented species of the <strong>genus</strong><br />

Cercospora Fresen., although he clearly recognised differences in<br />

their morphology. In a series of papers by Deighton, as well as<br />

others such as Sutton, Braun <strong>and</strong> Crous, the <strong>genus</strong> Cercospora<br />

was delimited based on its type species, Cercospora penicillata<br />

(Ces.) Fresen., while taxa formerly included in the <strong>genus</strong> by Chupp<br />

(1954) but differing in conidiophore arrangement, conidiogenesis,<br />

pigmentation, conidial catenulation, septation, <strong>and</strong> scar/hilum<br />

structure were allocated to other genera. Similar studies in which<br />

the type species were recollected <strong>and</strong> subjected to DNA sequence


Crous et al.<br />

Table 1. Isolates for which new sequences were generated.<br />

Anamorph Teleomorph Accession number 1 Host Country Collector GenBank Accession number<br />

Batcheloromyces eucalypti <strong>CBS</strong> 313.76; CPC 3632 Eucalyptus tessellaris Australia J.L. Alcorn EU019245<br />

Batcheloromyces leucadendri <strong>CBS</strong> 110892; CPC 1837 Leucadendron sp. South Africa L. Swart EU019246<br />

Batcheloromyces proteae <strong>CBS</strong> 110696; CPC 1518 Protea cynaroides South Africa L. Viljoen EU019247<br />

Capnobotryella renispora <strong>CBS</strong> 214.90*; <strong>CBS</strong> 176.88; IAM 13014; JCM 6932 Capnobotrys neessii Japan J. Sugiyama EU019248<br />

Catenulostroma abietis <strong>CBS</strong> 290.90 Man, skin lesion Netherl<strong>and</strong>s R.G.F. Wintermans EU019249<br />

Catenulostroma castellanii <strong>CBS</strong> 105.75*; ATCC 24788 Man, tinea nigra Venezuela — EU019250<br />

Catenulostroma chromoblastomycosum <strong>CBS</strong> 597.97 Man, chromoblastomycosis Zaire V. de Brouwere EU019251<br />

Catenulostroma elginense <strong>CBS</strong> 111030; CPC 1958 Protea gr<strong>and</strong>iceps South Africa J.E. Taylor EU019252<br />

Catenulostroma germanicum <strong>CBS</strong> 539.88 Stone Germany — EU019253<br />

Catenulostroma macowanii <strong>CBS</strong> 110756; CPC 1872 Protea nitida South Africa J.E. Taylor EU019254<br />

Catenulostroma microsporum Teratosphaeria microspora <strong>CBS</strong> 110890; CPC 1832 Protea cynaroides South Africa L. Swart EU019255<br />

Catenulostroma sp. Teratosphaeria<br />

<strong>CBS</strong> 118911; CPC 12085 Eucalyptus sp. Uruguay M.J. Wingfield EU019256<br />

pseudosuberosa<br />

Cercosporella centaureicola <strong>CBS</strong> 120253 Centaurea solstitiales Greece D. Berner EU019257<br />

Cibiessia dimorphospora <strong>CBS</strong> 120034; CPC 12636 Eucalyptus nitens Australia — EU019258<br />

Cibiessia minutispora CPC 13071* Eucalyptus henryii Australia A.J. Carnegie EU019259<br />

Cibiessia nontingens Teratosphaeria sp. <strong>CBS</strong> 120725*; CPC 13217 Eucalyptus tereticornis Australia B. Summerell EU019260<br />

<strong>Cladosporium</strong> bruhnei Davidiella allicina <strong>CBS</strong> 115683; ATCC 66670; CPC 5101 CCA-treated Douglas-fire pole U.S.A., New York C.J. Wang EU019261<br />

<strong>Cladosporium</strong> cladosporioides <strong>CBS</strong> 109.21; ATCC 11277; ATCC 200940; IFO 6368; IMI 049625 Sooty mould on Hedera helix U.K. — EU019262<br />

<strong>Cladosporium</strong> sphaerospermum <strong>CBS</strong> 188.54; ATCC 11290; IMI 049638 — — — EU019263<br />

<strong>Cladosporium</strong> uredinicola ATCC 46649 Hyperparasite on Cronartium<br />

fusiforme f. sp. quercum<br />

U.S.A., Alabama — EU019264<br />

Coccodinium bartschii <strong>CBS</strong> 121708; CPC 13861–13863 Sooty mould on unidentified tree Canada K.A. Seifert EU019265<br />

Dissoconium aciculare <strong>CBS</strong> 342.82*; CPC 1534 Erysiphe, on Medicago lupulina Germany T. Hijwegen EU019266<br />

Dissoconium commune “Mycosphaerella” communis <strong>CBS</strong> 114238*; CPC 10440 Eucalyptus globulus Spain J.P.M. Vazquez EU019267<br />

Dissoconium dekkeri “Mycosphaerella” lateralis <strong>CBS</strong> 567.89*; CPC 1535 Juniperus chinensis Netherl<strong>and</strong>s T. Hijwegen EU019268<br />

Fumagospora capnodioides Capnodium salicinum <strong>CBS</strong> 131.34 Sooty mould on Bursaria spinosa Indonesia — EU019269<br />

Hortaea werneckii <strong>CBS</strong> 107.67* Man, tinea nigra Portugal — EU019270<br />

Nothostrasseria dendritica Teratosphaeria dendritica CPC 12820 Eucalyptus nitens Australia A.J. Carnegie EU019271<br />

“Passalora” zambiae <strong>CBS</strong> 112970*; CPC 1228 Eucalyptus globulus Zambia T. Coutinho EU019272<br />

<strong>CBS</strong> 112971*; CMW 14782; CPC 1227 Eucalyptus globulus Zambia T. Coutinho EU019273<br />

Penidiella columbiana <strong>CBS</strong> 486.80 Paepalanthus columbianus Colombia W. Gams EU019274


Phylogenetic lineages in the Capnodiales<br />

Anamorph Teleomorph Accession number 1 Host Country Collector GenBank Accession number<br />

Penidiella nect<strong>and</strong>rae <strong>CBS</strong> 734.87*; ATCC 200932; INIFAT 87/45 Nect<strong>and</strong>ra coriacea Cuba R.F. Castañeda & EU019275<br />

G. Arnold<br />

Penidiella rigidophora <strong>CBS</strong> 314.95* Leaf litter of Smilax sp. Cuba R.F. Castañeda EU019276<br />

Penidiella strumelloidea <strong>CBS</strong> 114484*; VKM F-2534 Carex leaf, from stagnant water Russia S. Ozerskaya EU019277<br />

Penidiella venezuelensis <strong>CBS</strong> 106.75* Man, tinea nigra Venezuela D. Borelli EU019278<br />

Phaeotheca triangularis <strong>CBS</strong> 471.90* Wet surface of humidifier of<br />

airconditioning<br />

Belgium H. Beguin EU019279<br />

Phaeothecoidea eucalypti CPC 13010 Corymbia henryii Australia B. Summerell EU019280<br />

CPC 12918* Eucalyptus botryoides Australia B. Summerell EU019281<br />

Pleurophoma sp. Teratosphaeria fibrillosa CPC 1876 Protea nitida South Africa J.E. Taylor EU019282<br />

Pseudotaeniolina globosa <strong>CBS</strong> 109889* Rock Italy C. Urzi EU019283<br />

Ramularia pratensis var. pratensis CPC 11294 Rumex crispus Korea H.D. Shin EU019284<br />

Ramularia sp. <strong>CBS</strong> 324.87 On Mycosphaerella sp., leaf spot<br />

on Brassica sp.<br />

Netherl<strong>and</strong>s — EU019285<br />

Readeriella brunneotingens CPC 13303 Eucalyptus tereticornis Australia P.W. Crous EU019286<br />

Readeriella destructans <strong>CBS</strong> 111369*; CPC 1366 Eucalyptus gr<strong>and</strong>is Indonesia M.J. Wingfield EU019287<br />

Readeriella epicoccoides Teratosphaeria suttonii CPC 12352 Eucalyptus sp. U.S.A.,Hawaii W. Gams EU019288<br />

Readeriella eucalypti CPC 11186 Eucalyptus globulus Spain M.J. Wingfield EU019289<br />

Readeriella gauchensis <strong>CBS</strong> 120303*; CMW 17331 Eucalyptus gr<strong>and</strong>is Uruguay M.J. Wingfield EU019290<br />

Readeriella mirabilis <strong>CBS</strong> 116293; CPC 10506 Eucalyptus fastigata New Zeal<strong>and</strong> W. Gams EU019291<br />

Readeriella molleriana Teratosphaeria molleriana <strong>CBS</strong> 111164*; CMW 4940; CPC 1214 Eucalyptus globulus Portugal M.J. Wingfield EU019292<br />

Readeriella ovata complex CPC 18 Eucalyptus cladocalyx South Africa P.W. Crous EU019293<br />

<strong>CBS</strong> 111149; CPC 23 Eucalyptus cladocalyx South Africa P.W. Crous EU019294<br />

Readeriella stellenboschiana <strong>CBS</strong> 116428; CPC 10886 Eucalyptus sp. South Africa P.W. Crous EU019295<br />

Readeriella zuluensis <strong>CBS</strong> 120301*; CMW 17321 Eucalyptus gr<strong>and</strong>is South Africa M.J. Wingfield EU019296<br />

Septoria tritici Mycosphaerella graminicola <strong>CBS</strong> 100335; IPO 69001.61 Triticum aestivum — G.H.J. Kema EU019297<br />

<strong>CBS</strong> 110744; CPC 658 Triticum sp. South Africa P.W. Crous EU019298<br />

Trimmatostroma betulinum <strong>CBS</strong> 282.74 Betula verrucosa Netherl<strong>and</strong>s W.M. Loerakker EU019299<br />

Trimmatostroma salicis CPC 13571 Salix alba Germany U. Braun EU019300<br />

Teratosphaeria bellula <strong>CBS</strong> 111700; CPC 1821 Protea eximia South Africa J.E. Taylor EU019301<br />

Teratosphaeria mexicana CPC 12349 Eucalyptus sp. U.S.A.,Hawaii W. Gams EU019302<br />

Teratosphaeria nubilosa <strong>CBS</strong> 114419; CPC 10497 Eucalyptus globulus New Zeal<strong>and</strong> — EU019303<br />

<strong>CBS</strong> 116005*; CMW 3282; CPC 937 Eucalyptus globulus Australia A. Carnegie EU019304<br />

www.studiesinmycology.org


Crous et al.<br />

Table 1. (Continued).<br />

Anamorph Teleomorph Accession number 1 Host Country Collector GenBank Accession number<br />

Teratosphaeria ohnowa <strong>CBS</strong> 112896*; CMW 4937; CPC 1004 Eucalyptus gr<strong>and</strong>is South Africa M.J. Wingfield EU019305<br />

Teratosphaeria secundaria <strong>CBS</strong> 115608; CPC 504 Eucalyptus gr<strong>and</strong>is Brazil A.C. Alfenas EU019306<br />

Teratosphaeria sp. <strong>CBS</strong> 208.94; CPC 727 Eucalyptus gr<strong>and</strong>is Indonesia A.C. Alfenas EU019307<br />

1 ATCC: American Type Culture Collection, Virginia, U.S.A.; <strong>CBS</strong>: Centraalbureau voor Schimmelcultures, Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; CPC: Culture collection of Pedro Crous, housed at <strong>CBS</strong>; CMW: Culture collection of Mike Wingfield, housed at FABI,<br />

Pretoria, South Africa; IAM: Institute of Applied Microbiology, University of Tokyo, Institute of molecular <strong>and</strong> cellular bioscience, Tokyo, Japan; IFO: Institute For Fermentation, Osaka, Japan; IMI: International Mycological Institute, CABI-Bioscience,<br />

Egham, Bakeham Lane, U.K.; INIFAT: Alex<strong>and</strong>er Humboldt Institute for Basic Research in Tropical Agriculture, Ciudad de La Habana, Cuba; JCM: Japan Collection Of Microorganisms, RIKEN BioResource Center, Japan; VKM: All-Russian Collection of<br />

Microorganisms, Institute of Biochemistry <strong>and</strong> Physiology of Microorganisms, Russian Academy of Sciences, Pushchino, Russia.<br />

*Ex-type cultures.<br />

analysis were undertaken to characterise Mycosphaerella (Verkley<br />

et al. 2004), <strong>and</strong> anamorph genera such as Pseudocercospora<br />

Speg., Stigmina Sacc., Phaeoisariopsis Ferraris (Crous et al.<br />

2006a), Ramulispora Miura (Crous et al. 2003), Batcheloromyces<br />

Marasas, P.S. van Wyk & Knox-Dav. (Taylor et al. 2003),<br />

Phaeophleospora Rangel <strong>and</strong> Dothistroma Hulbary (Crous et al.<br />

2000, 2001, Barnes et al. 2004).<br />

To assess the phylogeny of the species selected for the present<br />

study, DNA sequences were generated of the 28S rRNA (LSU)<br />

gene. In a further attempt to address monophyletic groups within<br />

this complex, these data were integrated with their morphological<br />

characteristics. To further resolve pleomorphism among the species<br />

studied, isolates were examined on a range of cultural media to<br />

induce possible synanamorphs.<br />

Materials <strong>and</strong> methods<br />

Isolates<br />

Chosen isolates represent various species previously observed to<br />

be morphologically distinct from Mycosphaerella s. str. (Crous 1998,<br />

Crous et al. 2004a, b, 2006a, b, 2007b). In a few cases, specifically<br />

Teratosphaeria fibrillosa Syd. & P. Syd. <strong>and</strong> Coccodinium bartschii<br />

A. Massal., fresh material had to be collected from South Africa<br />

<strong>and</strong> Canada, respectively. Excised tissue pieces bearing ascomata<br />

were soaked in water for approximately 2 h, after which they were<br />

placed in the bottom of Petri dish lids, with the top half of the<br />

dish containing 2 % malt extract agar (MEA) (Gams et al. 2007).<br />

Ascospore germination patterns were examined after 24 h, <strong>and</strong><br />

single-ascospore <strong>and</strong> conidial cultures established as described by<br />

Crous (1998). Colonies were sub-cultured onto synthetic nutrientpoor<br />

agar (SNA), potato-dextrose agar (PDA), oatmeal agar (OA),<br />

MEA (Gams et al. 2007), <strong>and</strong> incubated at 25 °C under continuous<br />

near-ultraviolet light to promote sporulation.<br />

DNA phylogeny<br />

Fungal colonies were established on agar plates, <strong>and</strong> genomic<br />

DNA was isolated following the CTAB-based protocol described<br />

in Gams et al. (2007). <strong>The</strong> primers V9G (de Hoog & Gerrits van<br />

den Ende 1998) <strong>and</strong> LR5 (Vilgalys & Hester 1990) were used to<br />

amplify part of the nuclear rDNA operon spanning the 3’ end of<br />

the 18S rRNA gene (SSU), the first internal transcribed spacer<br />

(ITS1), the 5.8S rRNA gene, the second ITS region (ITS2) <strong>and</strong> the<br />

5’ end of the 28S rRNA gene (LSU). <strong>The</strong> primers ITS4 (White et<br />

al. 1990), LR0R (Rehner & Samuels 1994), LR3R (www.biology.<br />

duke.edu/fungi/mycolab/primers.htm), <strong>and</strong> LR16 (Moncalvo et al.<br />

1993), were used as internal sequence primers to ensure good<br />

quality sequences over the entire length of the amplicon. <strong>The</strong><br />

ITS1, ITS2 <strong>and</strong> 5.8S rRNA gene (ITS) were only sequenced for<br />

isolates of which these data were not available. <strong>The</strong> ITS data<br />

were not included in the analyses but deposited in GenBank<br />

where applicable. <strong>The</strong> PCR conditions, sequence alignment <strong>and</strong><br />

subsequent phylogenetic analysis using parsimony, distance <strong>and</strong><br />

Bayesian analyses followed the methods of Crous et al. (2006c).<br />

Gaps longer than 10 bases were coded as single events for the<br />

phylogenetic analyses; the remaining gaps were treated as new<br />

character states. Sequence data were deposited in GenBank<br />

(Table 1) <strong>and</strong> alignments in TreeBASE (www.treebase.org).<br />

Taxonomy<br />

Wherever possible, 30 measurements (× 1 000 magnification)


Phylogenetic lineages in the Capnodiales<br />

were made of structures mounted in lactic acid, with the extremes<br />

of spore measurements given in parentheses. Ascospores were<br />

frequently also mounted in water to observe mucoid appendages<br />

<strong>and</strong> sheaths. Colony colours (surface <strong>and</strong> reverse) were assessed<br />

after 1–2 mo on MEA at 25 °C in the dark, using the colour charts of<br />

Rayner (1970). All cultures obtained in this study are maintained in<br />

the culture collection of the Centraalbureau voor Schimmelcultures<br />

(<strong>CBS</strong>) in Utrecht, the Netherl<strong>and</strong>s (Table 1). Nomenclatural<br />

novelties <strong>and</strong> descriptions were deposited in MycoBank (www.<br />

MycoBank.org).<br />

Results<br />

DNA phylogeny<br />

Amplification products of approximately 1 700 bases were obtained<br />

for the isolates listed in Table 1. <strong>The</strong> LSU region of the sequences<br />

was used to obtain additional sequences from GenBank which<br />

were added to the alignment. <strong>The</strong> manually adjusted alignment<br />

contained 97 sequences (including the two outgroup sequences)<br />

<strong>and</strong> 844 characters including alignment gaps. Of the 844 characters<br />

used in the phylogenetic analysis, 308 were parsimony-informative,<br />

105 were variable <strong>and</strong> parsimony-uninformative, <strong>and</strong> 431 were<br />

constant.<br />

<strong>The</strong> parsimony analysis of the LSU region yielded 1 135 equally<br />

most parsimonious trees (TL = 1 502 steps; CI = 0.446; RI = 0.787;<br />

RC = 0.351), one of which is shown in Fig. 1. Three orders are<br />

represented by the ingroup isolates, namely Chaetothyriales (100<br />

% bootstrap support), Helotiales (100 % bootstrap support) <strong>and</strong><br />

Capnodiales (100 % bootstrap support). <strong>The</strong>se are discussed in<br />

detail in the Taxonomy <strong>and</strong> Discussion sections. A new collection of<br />

Coccodinium bartschii A. Massal clusters (100 % bootstrap support)<br />

with members of the Herpotrichiellaceae (Chaetothyriales), whereas<br />

the type species of the <strong>genus</strong> Trimmatostroma Corda, namely<br />

T. salicis Corda, as well as T. betulinum (Corda) S. Hughes, are<br />

allied (99 % bootstrap support) with the Dermateaceae (Helotiales).<br />

<strong>The</strong> Capnodiales encompasses members of the Capnodiaceae,<br />

Trichosphaeriaceae, Davidiellaceae, Schizothyriaceae <strong>and</strong> taxa<br />

traditionally placed in the Mycosphaerellaceae, which is divided<br />

here into the Teratosphaeriaceae, (65 % bootstrap support), <strong>and</strong><br />

the Mycosphaerellaceae (76 % bootstrap support), which contains<br />

several subclades. Also included in the Capnodiales are Devriesia<br />

staurophora (W.B. Kendr.) Seifert & N.L. Nick., Staninwardia<br />

suttonii Crous & Summerell <strong>and</strong> Capnobotryella renispora Sugiy.<br />

as sister taxa to Teratosphaeriaceae s. str. Neighbour-joining<br />

analysis using three substitution models on the sequence data<br />

yielded trees supporting the same topologies, but differed from<br />

the parsimony tree presented with regard to the order of the<br />

families <strong>and</strong> orders at the deeper nodes, e.g., the Helotiales <strong>and</strong><br />

Chaetothyriales are swapped around, as are the Capnodiaceae <strong>and</strong><br />

the Trichosphaeriaceae / Davidiellaceae (data not shown). Using<br />

neighbour-joining analyses, the Mycosphaerellaceae s. str. clade<br />

obtained 71 %, 70 % <strong>and</strong> 70 % bootstrap support respectively with<br />

the uncorrected “p”, Kimura 2-parameter <strong>and</strong> HKY85 substitution<br />

models wherease the Teratosphaeriaceae clade obtained 74 %, 79<br />

% <strong>and</strong> 78 % bootstrap support respectively with the same models.<br />

<strong>The</strong> Schizothyriaceae clade appeared basal in the Capnodiales,<br />

irrespective of which substitution model was used.<br />

Bayesian analysis was conducted on the same aligned LSU<br />

dataset using a general time-reversible (GTR) substitution model<br />

with inverse gamma rates <strong>and</strong> dirichlet base frequencies. <strong>The</strong><br />

Markov Chain Monte Carlo (MCMC) analysis of 4 chains started<br />

from a r<strong>and</strong>om tree topology <strong>and</strong> lasted 23 881 500 generations.<br />

Trees were saved each 100 generations, resulting in 238 815 saved<br />

trees. Burn-in was set at 22 000 000 generations after which the<br />

likelihood values were stationary, leaving 18 815 trees from which<br />

the consensus tree (Fig. 2) <strong>and</strong> posterior probabilities (PP’s) were<br />

calculated. <strong>The</strong> average st<strong>and</strong>ard deviation of split frequencies<br />

was 0.011508 at the end of the run. <strong>The</strong> same overall topology as<br />

that observed using parsimony was obtained, with the exception of<br />

the inclusion of Staninwardia suttonii in the Mycosphaerellaceae<br />

(PP value of 0.74) <strong>and</strong> not in the Teratosphaeriaceae. <strong>The</strong><br />

Mycosphaerellaceae s. str. clade, as well as the Teratosphaeriaceae<br />

clade, obtained a PP value of 1.00.<br />

Taxonomy<br />

Based on the dataset generated in this study, several well-supported<br />

genera could be distinguished in the Mycosphaerella complex (Figs<br />

1–2), for which we have identified morphological characters. <strong>The</strong>se<br />

genera, <strong>and</strong> a selection of their species, are treated below.<br />

Key to Mycosphaerella, <strong>and</strong> Mycosphaerella-like genera treated<br />

1. Ascomata thyrothecial; anamorph Zygosporium ................................................................................................................. Schizothyrium<br />

1. Ascomata pseudothecial ............................................................................................................................................................................ 2<br />

2. Ascospores with irregular, angular lumens typical of Davidiella; anamorph <strong>Cladosporium</strong> s. str. .............................................. Davidiella<br />

2. Ascospores guttulate or not, lacking angular lumens; anamorph other than <strong>Cladosporium</strong> ...................................................................... 3<br />

3. Ascomata frequently linked by superficial stroma; hamathecial tissue, ascospore sheath, multi-layered endotunica, prominent<br />

periphysoids, <strong>and</strong> ascospores turning brown in asci frequently observed ......................................................................... Teratosphaeria<br />

3. Ascomata not linked by superficial stroma; hamathecial tissue, ascospore sheath, multi-layered endotunica, prominent periphysoids,<br />

ascospores turning brown in asci not observed ......................................................................................................................................... 4<br />

4. Conidiophores solitary, pale brown, giving rise to primary <strong>and</strong> secondary, actively discharged conidia; anamorph Dissoconium<br />

.................................................................................................................................................................... teleomorph Mycosphaerella-like<br />

4. Conidiomata variable from solitary conidiophores to sporodochia, fascicles to pycnidia, but conidia not actively discharged .....................<br />

.................................................................................................................................................................................. Mycosphaerella s. str.<br />

www.studiesinmycology.org


Crous et al.<br />

100<br />

10 changes<br />

65<br />

100<br />

Athelia epiphylla AY586633<br />

Paullicorticium ansatum AY586693<br />

65<br />

63<br />

68<br />

100<br />

99<br />

100<br />

91<br />

67<br />

94<br />

Coccodinium bartschii CPC 13861<br />

Exophiala oligosperma AF050289<br />

Exophiala jeanselmei AF050271<br />

Fonsecaea pedrosoi AF050276<br />

Capronia mansonii AY004338<br />

Exophiala dermatitidis DQ823100<br />

Vibrissea flavovirens AY789426<br />

Vibrissea truncorum AY789402<br />

Trimmatostroma salicis CPC 13571<br />

Mollisia cinerea DQ470942<br />

Trimmatostroma betulinum <strong>CBS</strong> 282.74<br />

Phaeotheca triangularis <strong>CBS</strong> 471.90<br />

Capnodium coffeae DQ247800<br />

Capnodium salicinum <strong>CBS</strong> 131.34<br />

Scorias spongiosa DQ678075<br />

63<br />

68<br />

73<br />

92<br />

65<br />

98<br />

100<br />

56<br />

58<br />

100 87<br />

100<br />

100<br />

76<br />

Trichosphaeria pilosa AY590297 Trichosphaeriaceae<br />

<strong>Cladosporium</strong> cladosporioides <strong>CBS</strong> 109.21<br />

<strong>Cladosporium</strong> uredinicola ATCC 46649<br />

<strong>Cladosporium</strong> bruhnei <strong>CBS</strong> 115683<br />

<strong>Cladosporium</strong> sphaerospermum <strong>CBS</strong> 188.54<br />

64<br />

Schizothyrium pomi EF134947<br />

Schizothyrium pomi EF134948<br />

80 100<br />

100<br />

99<br />

100<br />

100<br />

100<br />

63<br />

64<br />

61<br />

100<br />

Herpotrichiellaceae,<br />

Chaetothyriales,<br />

Chaetothyriomycetes<br />

Vibrisseaceae<br />

Dermateaceae<br />

Capnodiaceae<br />

Schizothyriaceae<br />

“Passalora” zambiae <strong>CBS</strong> 112970<br />

“Passalora” zambiae <strong>CBS</strong> 112971<br />

Dissoconium aciculare <strong>CBS</strong> 342.82<br />

“Mycosphaerella” communis <strong>CBS</strong> 114238<br />

“Mycosphaerella” lateralis <strong>CBS</strong> 567.89<br />

Mycosphaerella graminicola <strong>CBS</strong> 100335<br />

Septoria tritici <strong>CBS</strong> 110744<br />

Cercosporella centaureicola <strong>CBS</strong> 120253<br />

Mycosphaerella punctiformis AY490776<br />

Ramularia sp. <strong>CBS</strong> 324.87<br />

Ramularia miae DQ885902<br />

Ramularia pratensis var. pratensis CPC 11294<br />

Capnobotryella renispora <strong>CBS</strong> 214.90<br />

100<br />

94<br />

88<br />

80<br />

99<br />

81<br />

Staninwardia suttonii DQ923535<br />

Devriesia staurophora DQ008150<br />

Devriesia staurophora DQ008151<br />

Pseudotaeniolina globosa <strong>CBS</strong> 109889<br />

Batcheloromyces proteae <strong>CBS</strong> 110696<br />

Batcheloromyces leucadendri <strong>CBS</strong> 110892<br />

Teratosphaeria alistairii DQ885901<br />

Penidiella columbiana <strong>CBS</strong> 486.80<br />

Teratosphaeria sp. <strong>CBS</strong> 208.94<br />

Teratosphaeria ohnowa <strong>CBS</strong> 112896<br />

Teratosphaeria secundaria <strong>CBS</strong> 115608<br />

Teratosphaeria flexuosa DQ246232<br />

Hortaea werneckii <strong>CBS</strong> 107.67<br />

Catenulostroma castellanii <strong>CBS</strong> 105.75<br />

Teratosphaeria bellula <strong>CBS</strong> 111700<br />

Catenulostroma elginense <strong>CBS</strong> 111030<br />

Teratosphaeria parva DQ246240<br />

100 Teratosphaeria parva DQ246243<br />

Catenulostroma germanicum <strong>CBS</strong> 539.88<br />

100<br />

Catenulostroma microsporum <strong>CBS</strong> 110890<br />

Catenulostroma abietis <strong>CBS</strong> 290.90<br />

Catenulostroma chromoblastomycosum <strong>CBS</strong> 597.97<br />

Phaeothecoidea eucalypti CPC 13010<br />

Phaeothecoidea eucalypti CPC 12918<br />

Teratosphaeria suberosa DQ246235<br />

Teratosphaeria pseudosuberosa <strong>CBS</strong> 118911<br />

93<br />

Teratosphaeria mexicana CPC 12349<br />

97 Teratosphaeria mexicana DQ246237<br />

Teratosphaeria readeriellophora DQ246238<br />

100<br />

93 Readeriella eucalypti CPC 11186<br />

Cibiessia minutispora CPC 13071<br />

Cibiessia dimorphospora <strong>CBS</strong> 120034<br />

Readeriella mirabilis <strong>CBS</strong> 116293<br />

Readeriella novaezel<strong>and</strong>iae DQ246239<br />

57 Nothostrasseria dendritica CPC 12820<br />

Batcheloromyces eucalypti <strong>CBS</strong> 313.76<br />

99 Readeriella ovata complex <strong>CBS</strong> 111149<br />

Readeriella ovata complex CPC 18<br />

Readeriella destructans <strong>CBS</strong> 111369<br />

Readeriella pulcherrima DQ246224<br />

Readeriella brunneotingens CPC 13303<br />

Readeriella dimorpha DQ923528<br />

Teratosphaeria molleriana DQ246219<br />

Teratosphaeria molleriana DQ246221<br />

Teratosphaeria molleriana <strong>CBS</strong> 111164<br />

Teratosphaeria fibrillosa CPC 1876<br />

Teratosphaeria maxii DQ885899<br />

Catenulostroma macowanii <strong>CBS</strong> 110756<br />

Teratosphaeria nubilosa <strong>CBS</strong> 114419<br />

Teratosphaeria nubilosa <strong>CBS</strong> 116005<br />

Readeriella stellenboschiana <strong>CBS</strong> 116428<br />

Readeriella gauchensis <strong>CBS</strong> 120303<br />

Teratosphaeria cryptica DQ246222<br />

Readeriella considenianae DQ923527<br />

Cibiessia nontingens <strong>CBS</strong> 120725<br />

Readeriella zuluensis <strong>CBS</strong> 120301<br />

Readeriella blakelyi DQ923526<br />

Teratosphaeria toledana DQ246230<br />

100<br />

Teratosphaeria suttonii DQ246227<br />

Readeriella epicoccoides CPC 12352<br />

Helotiales, Leotiomycetes<br />

Davidiellaceae<br />

Mycosphaerellaceae<br />

Teratosphaeriaceae<br />

Capnodiales, Dothideomycetes<br />

Fig. 1. One of 1 135 equally most parsimonious trees obtained from a heuristic search with 100 r<strong>and</strong>om taxon additions of the LSU sequence alignment using PAUP v. 4.0b10.<br />

<strong>The</strong> scale bar shows 10 changes, <strong>and</strong> bootstrap support values from 1 000 replicates are shown at the nodes. Thickened lines indicate the strict consensus branches <strong>and</strong> extype<br />

sequences are printed in bold face. <strong>The</strong> tree was rooted to two sequences obtained from GenBank (Athelia epiphylla AY586633 <strong>and</strong> Paullicorticium ansatum AY586693).


Phylogenetic lineages in the Capnodiales<br />

1.00<br />

0.62<br />

0.1 expected changes per site<br />

Athelia epiphylla AY586633<br />

Paullicorticium ansatum AY586693<br />

0.62<br />

1.00<br />

0.75<br />

1.00<br />

1.00<br />

0.99<br />

0.99<br />

0.85<br />

0.97<br />

Capronia mansonii AY004338<br />

Exophiala dermatitidis DQ823100<br />

Fonsecaea pedrosoi AF050276<br />

Coccodinium bartschii CPC 13861<br />

Exophiala oligosperma AF050289<br />

Exophiala jeanselmei AF050271<br />

Vibrisseaceae<br />

0.97<br />

Vibrissea flavovirens AY789426<br />

Vibrissea truncorum AY789402<br />

1.00<br />

Trimmatostroma salicis CPC 13571<br />

Mollisia cinerea DQ470942<br />

Trimmatostroma betulinum <strong>CBS</strong> 282.74<br />

Phaeotheca triangularis <strong>CBS</strong> 471.90<br />

Capnodium coffeae DQ247800<br />

Capnodium salicinum <strong>CBS</strong> 131.34<br />

1.00<br />

1.00<br />

1.00<br />

0.80<br />

0.74<br />

1.00<br />

0.55<br />

1.00<br />

1.00<br />

0.91<br />

Dermateaceae<br />

Herpotrichiellaceae,<br />

Chaetothyriales,<br />

Chaetothyriomycetes<br />

Helotiales,<br />

Leotiomycetes<br />

Capnodiaceae<br />

Scorias spongiosa DQ678075<br />

1.00<br />

Trichosphaeria pilosa AY590297 Trichosphaeriaceae<br />

<strong>Cladosporium</strong> cladosporioides <strong>CBS</strong> 109.21<br />

<strong>Cladosporium</strong> uredinicola ATCC 46649<br />

1.00 <strong>Cladosporium</strong> bruhnei <strong>CBS</strong> 115683<br />

0.95 <strong>Cladosporium</strong> sphaerospermum <strong>CBS</strong> 188.54<br />

Staninwardia suttonii DQ923535<br />

1.00 Schizothyrium pomi EF134947<br />

Schizothyrium pomi EF134948<br />

Schizothyriaceae<br />

1.00<br />

1.00<br />

0.83 1.00<br />

0.72 1.00<br />

1.00 0.58<br />

0.79<br />

0.89 0.62 1.00<br />

0.99<br />

1.00<br />

0.58<br />

0.94<br />

0.56<br />

0.83<br />

0.94<br />

0.77<br />

0.51<br />

1.00<br />

1.00<br />

0.77<br />

0.85<br />

Dissoconium aciculare <strong>CBS</strong> 342.82<br />

“Mycosphaerella” communis <strong>CBS</strong> 114238<br />

“Mycosphaerella” lateralis <strong>CBS</strong> 567.89<br />

Passalora zambiae <strong>CBS</strong> 112970<br />

Passalora zambiae <strong>CBS</strong> 112971<br />

Cercosporella centaureicola <strong>CBS</strong> 120253<br />

Mycosphaerella punctiformis AY490776<br />

Ramularia sp. <strong>CBS</strong> 324.87<br />

Ramularia miae DQ885902<br />

Ramularia pratensis var. pratensis CPC 11294<br />

0.57 Mycosphaerella graminicola <strong>CBS</strong> 100335<br />

1.00 Septoria tritici <strong>CBS</strong> 110744<br />

Capnobotryella renispora <strong>CBS</strong> 214.90<br />

Pseudotaeniolina globosa <strong>CBS</strong> 109889<br />

Catenulostroma chromoblastomycosum <strong>CBS</strong> 597.97<br />

Teratosphaeria bellula <strong>CBS</strong> 111700<br />

Hortaea werneckii <strong>CBS</strong> 107.67<br />

Catenulostroma castellanii <strong>CBS</strong> 105.75<br />

Batcheloromyces proteae <strong>CBS</strong> 110696<br />

Batcheloromyces leucadendri <strong>CBS</strong> 110892<br />

Teratosphaeria alistairii DQ885901<br />

Penidiella columbiana <strong>CBS</strong> 486.80<br />

Teratosphaeria sp. <strong>CBS</strong> 208.94<br />

Teratosphaeria ohnowa <strong>CBS</strong> 112896<br />

Teratosphaeria secundaria <strong>CBS</strong> 115608<br />

Teratosphaeria flexuosa DQ246232<br />

Catenulostroma elginense <strong>CBS</strong> 111030<br />

Teratosphaeria parva DQ246240<br />

1.00<br />

1.00<br />

1.00<br />

1.00<br />

1.00<br />

1.00<br />

0.66<br />

0.76<br />

Teratosphaeria parva DQ246243<br />

Devriesia staurophora DQ008150<br />

Devriesia staurophora DQ008151<br />

Catenulostroma germanicum <strong>CBS</strong> 539.88<br />

Catenulostroma microsporum <strong>CBS</strong> 110890<br />

Catenulostroma abietis <strong>CBS</strong> 290.90<br />

Phaeothecoidea eucalypti CPC 13010<br />

Phaeothecoidea eucalypti CPC 12918<br />

Teratosphaeria pseudosuberosa <strong>CBS</strong> 118911<br />

Teratosphaeria suberosa DQ246235<br />

0.97<br />

Teratosphaeria mexicana CPC 12349<br />

Teratosphaeria mexicana <strong>CBS</strong> 110502<br />

Readeriella novaezel<strong>and</strong>iae DQ246239<br />

Nothostrasseria dendritica CPC 12820<br />

Readeriella mirabilis <strong>CBS</strong> 116293<br />

Cibiessia dimorphospora <strong>CBS</strong> 120034<br />

Cibiessia minutispora CPC 13071<br />

Teratosphaeria readeriellophora DQ246238<br />

Readeriella eucalypti CPC 11186<br />

1.00<br />

Readeriella brunneotingens CPC 13303<br />

Readeriella stellenboschiana <strong>CBS</strong> 116428<br />

Readeriella gauchensis <strong>CBS</strong> 120303<br />

Readeriella considenianae DQ923527<br />

Teratosphaeria cryptica DQ246222<br />

Cibiessia nontingens <strong>CBS</strong> 120725<br />

Readeriella zuluensis <strong>CBS</strong> 120301<br />

Teratosphaeria toledana DQ246230<br />

Teratosphaeria suttonii DQ246227<br />

Readeriella epicoccoides CPC 12352<br />

1.00<br />

Readeriella blakelyi DQ923526<br />

Readeriella ovata complex <strong>CBS</strong> 111149<br />

Readeriella ovata complex CPC 18<br />

Batcheloromyces eucalypti <strong>CBS</strong> 313.76<br />

Readeriella destructans <strong>CBS</strong> 111369<br />

Readeriella pulcherrima DQ246224<br />

Teratosphaeria nubilosa <strong>CBS</strong> 114419<br />

Teratosphaeria nubilosa <strong>CBS</strong> 116005<br />

Readeriella dimorpha DQ923528<br />

Teratosphaeria molleriana DQ246219<br />

Teratosphaeria molleriana DQ246221<br />

Teratosphaeria molleriana <strong>CBS</strong> 111164<br />

Teratosphaeria fibrillosa CPC 1876<br />

Catenulostroma macowanii <strong>CBS</strong> 110756<br />

Teratosphaeria maxii DQ885899<br />

Davidiellaceae<br />

Mycosphaerellaceae Teratosphaeriaceae<br />

Capnodiales, Dothideomycetes<br />

Fig. 2. Consensus phylogram (50 % majority rule) of 18 815 trees resulting from a Bayesian analysis of the LSU sequence alignment using MrBayes v. 3.1.2. Bayesian posterior<br />

probabilities are indicated at the nodes. Ex-type sequences are printed in bold face. <strong>The</strong> tree was rooted to two sequences obtained from GenBank (Athelia epiphylla AY586633<br />

<strong>and</strong> Paullicorticium ansatum AY586693).<br />

www.studiesinmycology.org


Crous et al.<br />

Treatment of phylogenetic clades<br />

Davidiellaceae clade<br />

Davidiella Crous & U. Braun, Mycol. Progr. 2: 8. 2003.<br />

Type species: Davidiella tassiana (De Not.) Crous & U. Braun,<br />

Mycol. Progr. 2: 8. 2003.<br />

Basionym: Sphaerella tassiana De Not., Sferiacei Italici 1: 87.<br />

1863.<br />

Description: Schubert et al. (2007 – this volume).<br />

Anamorph: <strong>Cladosporium</strong> Link, Ges. Naturf. Freunde Berlin Mag.<br />

Neuesten Entdeck. Gesammten Naturk. 7: 37. 1816.<br />

Type species: <strong>Cladosporium</strong> herbarum (Pers. : Fr.) Link, Ges.<br />

Naturf. Freunde Berlin Mag. Neuesten Entdeck. Gesammten<br />

Naturk. 7: 37. 1816.<br />

Basionym: Dematium herbarum Pers., Ann. Bot. (Usteri), 11 Stück:<br />

32. 1794: Fr., Syst. Mycol. 3: 370. 1832.<br />

Description: Schubert et al. (2007 – this volume).<br />

Notes: <strong>The</strong> <strong>genus</strong> Davidiella (Davidiellaceae) was recently<br />

introduced for teleomorphs of <strong>Cladosporium</strong> s. str. (Braun et al.<br />

2003). <strong>The</strong> <strong>genus</strong> <strong>Cladosporium</strong> is well-established, <strong>and</strong> contains<br />

around 772 names (Dugan et al. 2004), while Davidiella presently<br />

has 33 names (www.MycoBank.org), of which only around five<br />

have acknowledged <strong>Cladosporium</strong> states.<br />

Teratosphaeriaceae clade<br />

Teratosphaeria Syd. & P. Syd., Ann. Mycol. 10: 39. 1912.<br />

Type species: Teratosphaeria fibrillosa Syd. & P. Syd., Ann. Mycol.<br />

10: 40. 1912. Fig. 3.<br />

Description: Crous et al. (2004a; figs 182–185).<br />

Notes: Although <strong>similar</strong> in morphology, the <strong>genus</strong> Teratosphaeria<br />

was separated from Mycosphaerella based on its ascomatal<br />

arrangement, <strong>and</strong> periphysate ostioles (Müller & Oehrens 1982).<br />

It was later synonymised under Mycosphaerella by Taylor et<br />

al. (2003), who showed that the type species clustered within<br />

Mycosphaerella based on ITS DNA sequence data. <strong>The</strong> LSU<br />

sequence data generated in the present study, has clearly shown<br />

that Mycosphaerella is polyphyletic, thus contradicting earlier<br />

reports of monophyly by Crous et al. (2000) <strong>and</strong> Goodwin et al.<br />

(2001), which were based on ITS data.<br />

A re-examination of T. fibrillosa, the type species of<br />

Teratosphaeria, revealed several morphological features that<br />

characterise the majority of the taxa clustering in the clade, though<br />

several characters have been lost in some of the small-spored<br />

species. <strong>The</strong>se characters are discussed below:<br />

1. Teratosphaeria fibrillosa has a superficial stroma linking<br />

ascomata together, almost appearing like a spider’s web on the leaf<br />

surface. Although this feature is not seen in other taxa in this clade,<br />

some species, such as M. suberosa Crous, F.A. Ferreira, Alfenas<br />

& M.J. Wingf. <strong>and</strong> M. pseudosuberosa Crous & M.J. Wingf. have<br />

a superficial stroma, into which the ascomata are inbedded (Crous<br />

1998, Crous et al. 2006b).<br />

2. Ascospores of Teratosphaeria become brown <strong>and</strong><br />

verruculose while still in their asci. This feature is commonly<br />

observed in species such as M. jonkershoekensis P.S. van Wyk,<br />

Marasas & Knox-Dav., M. alistairii Crous, M. mexicana Crous, M.<br />

maxii Crous <strong>and</strong> M. excentricum Crous & Carnegie (Crous 1998,<br />

Crous & Groenewald 2006a, b, Crous et al. 2007b).<br />

3. A few ascomata of T. fibrillosa were found to have<br />

some pseudoparaphysoidal remnants (cells to distinguish<br />

pseudoparaphyses), though they mostly disappear with age. This<br />

feature is rather uncommon, though pseudoparaphyses were<br />

observed in ascomata of M. eucalypti (Wakef.) Hansf.<br />

4. Ascospores of Teratosphaeria were found to be covered in<br />

a mucous sheath, which is commonly observed in other taxa in this<br />

clade, such as M. bellula Crous & M.J. Wingf., M. pseudocryptica<br />

Crous, M. suberosa, M. pseudosuberosa, M. associata Crous &<br />

Carnegie, M. dendritica Crous & Summerell <strong>and</strong> M. fimbriata Crous<br />

& Summerell (Crous et al. 2004b, 2006b, 2007b). Re-examination<br />

of fresh collections also revealed ascospores of M. cryptica (Cooke)<br />

Hansf. <strong>and</strong> M. nubilosa (Cooke) Hansf. to have a weakly definable<br />

sheath. Germinating ascospores of species in this clade all exhibit<br />

a prominent mucoid sheath.<br />

5. Asci of T. fibrillosa were observed to have a multi-layered<br />

endotunica, which, although not common, can be seen in species<br />

such as M. excentrica, M. maxii, M. alistairii, M. pseudosuberosa,<br />

M. fimbriata (Crous et al. 2006b, 2007b, Crous & Groenewald<br />

2006a, b), <strong>and</strong> also M. nubilosa.<br />

6. Finally, ascomata of T. fibrillosa <strong>and</strong> T. proteae-arboreae<br />

P.S. van Wyk, Marasas & Knox-Dav. have well-developed<br />

ostiolar periphyses, which are also present in species such as M.<br />

suberosa, M. pseudosuberosa, M. maxii <strong>and</strong> T. microspora Joanne<br />

E. Taylor & Crous (Crous 1998, Crous et al. 2004a, b, 2006b).<br />

Morphologically thus, the Teratosphaeria clade is distinguishable<br />

from Mycosphaerella s. str., though these differences are less<br />

pronounced in some of the smaller-spored species. Based on<br />

these distinct morphological features, as well as its phylogenetic<br />

position within the Capnodiales, a new family is herewith proposed<br />

to accommodate species of Teratosphaeria:<br />

Teratosphaeriaceae Crous & U. Braun, fam. nov. MycoBank<br />

MB504464.<br />

Ascomata pseudotheciales, superficiales vel immersa, saepe in stromate ex cellulis<br />

brunneis pseudoparenchymatibus disposita, globulares, uniloculares, papillata,<br />

apice ostiolato, periphysata, saepe cum periphysoidibus; tunica multistratosa, ex<br />

cellulis brunneis angularibus composita, strato interiore ex cellulis applanatis hyalinis;<br />

saepe cum pseudoparaphysibus subcylindricis, ramosis, septatis, anastomosibus.<br />

Asci fasciculati, octospori, bitunicati, saepe cum endotunica multistratosa.<br />

Ascosporae ellipsoideae-fusiformes vel obovoideae, 1-septatae, hyalinae, deinde<br />

pallide brunneae et verruculosae, saepe mucosae.<br />

Ascomata pseudothecial, superficial to immersed, frequently<br />

situated in a stroma of brown pseudoparenchymatal cells,<br />

globose, unilocular, papillate, ostiolate, canal periphysate, with<br />

periphysoids frequently present; wall consisting of several layers<br />

of brown textura angularis; inner layer of flattened, hyaline cells.<br />

Pseudoparaphyses frequently present, subcylindrical, branched,<br />

septate, anastomosing. Asci fasciculate, 8-spored, bitunicate,<br />

frequently with multi-layered endotunica. Ascospores ellipsoidfusoid<br />

to obovoid, 1-septate, hyaline, but becoming pale brown <strong>and</strong><br />

verruculose, frequently covered in mucoid sheath.<br />

Typus: Teratosphaeria Syd. & P. Syd., Ann. Mycol. 10: 39. 1912.<br />

Teratosphaeria africana (Crous & M.J. Wingf.) Crous & U. Braun,<br />

comb. nov. MycoBank MB504466.


Phylogenetic lineages in the Capnodiales<br />

Fig. 3. Teratosphaeria fibrillosa (epitype material). A. Leaf spots. B. Subepidermal ascomata linked by means of stromatic tissue. C. Paraphyses among asci. D. Periphysoids. E.<br />

Ascospores becoming brown in asci. F–G. Multi-layered endotunica. H–K. Ascospores, becoming brown <strong>and</strong> verruculose. L–M. Germinating ascospores. Scale bars = 10 µm.<br />

Basionym: Mycosphaerella africana Crous & M.J. Wingf., Mycologia<br />

88: 450. 1996.<br />

Teratosphaeria associata (Crous & Carnegie) Crous & U. Braun,<br />

comb. nov. MycoBank MB504467.<br />

Basionym: Mycosphaerella associata Crous & Carnegie, Fungal<br />

Diversity 26: 159. 2007.<br />

Teratosphaeria alistairii (Crous) Crous & U. Braun, comb. nov.<br />

MycoBank MB504468.<br />

Basionym: Mycosphaerella alistairii Crous, in Crous & Groenewald,<br />

Fungal Planet, No. 4. 2006.<br />

Anamorph: Batcheloromyces sp.<br />

www.studiesinmycology.org


Crous et al.<br />

Teratosphaeria bellula (Crous & M.J. Wingf.) Crous & U. Braun,<br />

comb. nov. MycoBank MB504469.<br />

Basionym: Mycosphaerella bellula Crous & M.J. Wingf., Mycotaxon<br />

46: 20. 1993.<br />

Teratosphaeria cryptica (Cooke) Crous & U. Braun, comb. nov.<br />

MycoBank MB504470.<br />

Basionym: Sphaerella cryptica Cooke, Grevillea 20: 5. 1891.<br />

≡ Mycosphaerella cryptica (Cooke) Hansf., Proc. Linn. Soc. New South<br />

Wales 81: 35. 1956.<br />

Anamorph: Readeriella nubilosa (Ganap. & Corbin) Crous & U.<br />

Braun, comb. nov. MycoBank MB504471.<br />

Basionym: Colletogloeum nubilosum Ganap. & Corbin, Trans. Brit.<br />

Mycol. Soc. 72: 237. 1979.<br />

≡ Colletogloeopsis nubilosum (Ganap. & Corbin) Crous & M.J. Wingf.,<br />

Canad. J. Bot. 75: 668. 1997.<br />

Teratosphaeria dendritica (Crous & Summerell) Crous & U.<br />

Braun, comb. nov. MycoBank MB504472.<br />

Basionym: Mycosphaerella dendritica Crous & Summerell, Fungal<br />

Diversity 26: 161. 2007.<br />

Anamorph: Nothostrasseria dendritica (Hansf.) Nag Raj, Canad.<br />

J. Bot. 61: 25. 1983.<br />

Basionym: Spilomyces dendriticus Hansf., Proc. Linn. Soc. New<br />

South Wales 81: 32. 1956.<br />

Teratosphaeria excentrica (Crous & Carnegie) Crous & U. Braun,<br />

comb. nov. MycoBank MB504473.<br />

Basionym: Mycosphaerella excentrica Crous & Carnegie, Fungal<br />

Diversity 26: 164. 2007.<br />

Anamorph: Catenulostroma excentricum (B. Sutton & Ganap.)<br />

Crous & U. Braun, comb. nov. MycoBank MB504475.<br />

Basionym: Trimmatostroma excentricum B. Sutton & Ganap., New<br />

Zeal<strong>and</strong> J. Bot. 16: 529. 1978.<br />

Teratosphaeria fibrillosa Syd. & P. Syd., Ann. Mycol. 10: 40.<br />

1912.<br />

≡ Mycosphaerella fibrillosa (Syd. & P. Syd.) Joanne E. Taylor & Crous,<br />

Mycol. Res. 107: 657. 2003.<br />

Specimens examined: South Africa, Western Cape Province, Bains Kloof near<br />

Wellington, on living leaves of Protea gr<strong>and</strong>iflora, 26 Feb. 1911, E.M. Doidge,<br />

holotype PREM; Stellenbosch, Jonkershoek valley, S33° 59’ 44.7”, E18° 58’ 50.6”,<br />

1 Apr. 2007, on leaves of Protea sp., P.W. Crous & L. Mostert, epitype designated<br />

here <strong>CBS</strong> H-19913, culture ex-epitype <strong>CBS</strong> 121707 = CPC 13960.<br />

Teratosphaeria fimbriata (Crous & Summerell) Crous & U. Braun,<br />

comb. nov. MycoBank MB504476.<br />

Basionym: Mycosphaerella fimbriata Crous & Summerell, Fungal<br />

Diversity 26: 166. 2007.<br />

Teratosphaeria flexuosa (Crous & M.J. Wingf.) Crous & U. Braun,<br />

comb. nov. MycoBank MB504477.<br />

Basionym: Mycosphaerella flexuosa Crous & M.J. Wingf., Mycol.<br />

Mem. 21: 58. 1998.<br />

Teratosphaeria gamsii (Crous) Crous & U. Braun, comb. nov.<br />

MycoBank MB504478.<br />

Basionym: Mycosphaerella gamsii Crous, Stud. Mycol. 55: 113.<br />

2006.<br />

Teratosphaeria jonkershoekensis (P.S. van Wyk, Marasas &<br />

Knox-Dav.) Crous & U. Braun, comb. nov. MycoBank MB504479.<br />

Basionym: Mycosphaerella jonkershoekensis P.S. van Wyk,<br />

Marasas & Knox-Dav., J. S. African Bot. 41: 234. 1975.<br />

Teratosphaeria maxii (Crous) Crous & U. Braun, comb. nov.<br />

MycoBank MB504480.<br />

Basionym: Mycosphaerella maxii Crous, in Crous & Groenewald,<br />

Fungal Planet No. 6. 2006.<br />

Teratosphaeria mexicana (Crous) Crous & U. Braun, comb. nov.<br />

MycoBank MB504481.<br />

Basionym: Mycosphaerella mexicana Crous, Mycol. Mem. 21: 81.<br />

1998.<br />

Teratosphaeria microspora Joanne E. Taylor & Crous, Mycol.<br />

Res. 104: 631. 2000.<br />

≡ Mycosphaerella microspora (Joanne E. Taylor & Crous) Joanne E.<br />

Taylor & Crous, Mycol. Res. 107: 657. 2003.<br />

Anamorph: Catenulostroma microsporum (Joanne E. Taylor &<br />

Crous) Crous & U. Braun, comb. nov. MycoBank MB504482.<br />

Basionym: Trimmatostroma microsporum Joanne E. Taylor &<br />

Crous, Mycol. Res. 104: 631. 2000.<br />

Teratosphaeria molleriana (Thüm.) Crous & U. Braun, comb.<br />

nov. MycoBank MB504483.<br />

Basionym: Sphaerella molleriana Thüm., Revista Inst. Sci. Lit.<br />

Coimbra 28: 31. 1881.<br />

≡ Mycosphaerella molleriana (Thüm) Lindau, Nat. Pfanzenfam. 1: 424.<br />

1897.<br />

= Mycosphaerella vespa Carnegie & Keane, Mycol. Res. 102: 1275. 1998.<br />

= Mycosphaerella ambiphylla A. Maxwell, Mycol. Res. 107: 354. 2003.<br />

Anamorph: Readeriella molleriana (Crous & M.J. Wingf.) Crous &<br />

U. Braun, comb. nov. MycoBank MB504484.<br />

Basionym: Colletogloeopsis molleriana Crous & M.J. Wingf.,<br />

Canad. J. Bot. 75: 670. 1997.<br />

Teratosphaeria nubilosa (Cooke) Crous & U. Braun, comb. nov.<br />

MycoBank MB504485.<br />

Basionym: Sphaerella nubilosa Cooke, Grevillea 19: 61. 1892.<br />

≡ Mycosphaerella nubilosa (Cooke) Hansf., Proc. Linn. Soc. New South<br />

Wales 81: 36. 1965.<br />

= Mycosphaerella juvenis Crous & M.J. Wingf., Mycologia 88: 453. 1996.<br />

Teratosphaeria ohnowa (Crous & M.J. Wingf.) Crous & U. Braun,<br />

comb. nov. MycoBank MB504486.<br />

Basionym: Mycosphaerella ohnowa Crous & M.J. Wingf., Stud.<br />

Mycol. 50: 206. 2004.<br />

Teratosphaeria parkiiaffinis (Crous & M.J. Wingf.) Crous & U.<br />

Braun, comb. nov. MycoBank MB504487.<br />

Basionym: Mycosphaerella parkiiaffinis Crous & M.J. Wingf., Fungal<br />

Diversity 26: 168. 2007.<br />

Teratosphaeria parva (R.F. Park & Keane) Crous & U. Braun,<br />

comb. nov. MycoBank MB504488.<br />

Basionym: Mycosphaerella parva R.F. Park & Keane, Trans. Brit.<br />

Mycol. Soc. 79: 99. 1982.<br />

= Mycosphaerella gr<strong>and</strong>is Carnegie & Keane, Mycol. Res. 98: 414. 1994.<br />

Teratosphaeria perpendicularis (Crous & M.J. Wingf.) Crous & U.<br />

Braun, comb. nov. MycoBank MB504489.<br />

Basionym: Mycosphaerella perpendicularis Crous & M.J. Wingf.,<br />

Stud. Mycol. 55: 113. 2006.<br />

Teratosphaeria pluritubularis (Crous & Mansilla) Crous & U.<br />

Braun, comb. nov. MycoBank MB504490.<br />

Basionym: Mycosphaerella pluritubularis Crous & Mansilla, Stud.<br />

Mycol. 55: 114. 2006.<br />

10


Phylogenetic lineages in the Capnodiales<br />

Teratosphaeria pseudafricana (Crous & T.A. Cout.) Crous & U.<br />

Braun, comb. nov. MycoBank MB504491.<br />

Basionym: Mycosphaerella pseudafricana Crous & T.A. Cout.,<br />

Stud. Mycol. 55: 115. 2006.<br />

Teratosphaeria pseudocryptica (Crous) Crous & U. Braun, comb.<br />

nov. MycoBank MB504492.<br />

Basionym: Mycosphaerella pseudocryptica Crous, Stud. Mycol. 55:<br />

6. 2006.<br />

Anamorph: Readeriella sp.<br />

Teratosphaeria pseudosuberosa (Crous & M.J. Wingf.) Crous &<br />

U. Braun, comb. nov. MycoBank MB504493.<br />

Basionym: Mycosphaerella pseudosuberosa Crous & M.J. Wingf.,<br />

Stud. Mycol. 55: 118. 2006.<br />

Anamorph: Catenulostroma sp.<br />

Teratosphaeria quasicercospora (Crous & T.A. Cout.) Crous & U.<br />

Braun, comb. nov. MycoBank MB504494.<br />

Basionym: Mycosphaerella quasicercospora Crous & T.A. Cout.,<br />

Stud. Mycol. 55: 119. 2006.<br />

Teratosphaeria readeriellophora (Crous & Mansilla) Crous & U.<br />

Braun, comb. nov. MycoBank MB504495.<br />

Basionym: Mycosphaerella readeriellophora Crous & Mansilla,<br />

Stud. Mycol. 50: 207. 2004.<br />

Anamorph: Readeriella readeriellophora Crous & Mansilla, Stud.<br />

Mycol. 50: 207. 2004. Fig. 18.<br />

Teratosphaeria secundaria (Crous & Alfenas) Crous & U. Braun,<br />

comb. nov. MycoBank MB504496.<br />

Basionym: Mycosphaerella secundaria Crous & Alfenas, Stud.<br />

Mycol. 55: 122. 2006.<br />

Teratosphaeria stramenticola (Crous & Alfenas) Crous & U.<br />

Braun, comb. nov. MycoBank MB504497.<br />

Basionym: Mycosphaerella stramenticola Crous & Alfenas, Stud.<br />

Mycol. 55: 123. 2006.<br />

Teratosphaeria suberosa (Crous, F.A. Ferreira, Alfenas & M.J.<br />

Wingf.) Crous & U. Braun, comb. nov. MycoBank MB504498.<br />

Basionym: Mycosphaerella suberosa Crous, F.A. Ferreira, Alfenas<br />

& M.J. Wingf., Mycologia 85: 707. 1993.<br />

Teratosphaeria suttonii (Crous & M.J. Wingf.) Crous & U. Braun,<br />

comb. nov. MycoBank MB504499.<br />

Basionym: Mycosphaerella suttonii Crous & M.J. Wingf. (suttoniae),<br />

Canad. J. Bot. 75: 783. 1997.<br />

Anamorph: Readeriella epicoccoides (Cooke & Massee) Crous &<br />

U. Braun, comb. nov. MycoBank MB504500.<br />

Basionym: Cercospora epicoccoides Cooke & Massee apud Cooke,<br />

Grevillea 19: 91. 1891.<br />

≡ Phaeophleospora epicoccoides (Cooke & Massee) Crous, F.A. Ferreira<br />

& B. Sutton, S. African J. Bot. 63: 113. 1997.<br />

≡ Kirramyces epicoccoides (Cooke & Massee) J. Walker, B. Sutton &<br />

Pascoe, Mycol. Res. 96: 919. 1992.<br />

= Hendersonia gr<strong>and</strong>ispora McAlp., Proc. Linn. Soc. New South Wales 28:<br />

99. 1903.<br />

= Phaeoseptoria eucalypti Hansf., Proc. Linn. Soc. New South Wales 82: 225.<br />

1957.<br />

= Phaeoseptoria luzonensis T. Kobayashi, Trans. Mycol. Soc. Japan 19: 377.<br />

1978.<br />

Synanamorph: Pseudocercospora sp.<br />

Teratosphaeria toledana (Crous & Bills) Crous & U. Braun, comb.<br />

nov. MycoBank MB504501.<br />

Basionym: Mycosphaerella toledana Crous & Bills, Stud. Mycol. 50:<br />

208. 2004.<br />

Anamorph: Readeriella toledana (Crous & Bills) Crous & U. Braun,<br />

comb. nov. MycoBank MB504502.<br />

Basionym: Phaeophleospora toledana Crous & Bills, Stud. Mycol.<br />

50: 208. 2004.<br />

Key to treated anamorph genera of Teratosphaeria (Teratosphaeriaceae)<br />

1. Hyphae submerged to superficial, disarticulating into arthroconidia .......................................................................................................... 2<br />

1. Hyphae not disarticulating into arthroconidia ............................................................................................................................................. 3<br />

2. Mature, brown hyphae disarticulating into thick-walled, spherical, smooth to verruculose 0(–2) transversely septate,<br />

brown conidia ............................................................................................................................ Pseudotaeniolina (= Friedmanniomyces)<br />

2. Hyphae superficial, brown to green-brown, smooth, disarticulating to form pale brown, cylindrical, 0–3-septate conidia with subtruncate<br />

ends, frequently with a Readeriella synanamorph ....................................................................................................................... Cibiessia<br />

3. Hyphal ends forming endoconidia; hyphae pale to medium brown, verruculose, end cells dividing into several brown, verruculose,<br />

thick-walled, ellipsoid to obovoid endoconidia ................................................................................................................. Phaeothecoidea<br />

3. Endoconidia absent.................................................................................................................................................................................... 4<br />

4. Conidiogenous cells integrated in hyphae; well-developed conidiomata or long, solitary, macronematous, terminally penicillate<br />

conidiophores absent ................................................................................................................................................................................. 5<br />

4. Conidiomata well-developed or with long, solitary, terminally penicillate conidiophores ........................................................................... 7<br />

5. Conidia in chains, holoblastic, pseudocladosporium-like in morphology, but scars <strong>and</strong> hila not excessively thickened, nor refractive,<br />

producing chlamydospores in culture; species are mostly heat resistant .................................................................................... Devriesia<br />

5. Conidia solitary on indistict to well defined phialides on hyphae ............................................................................................................... 6<br />

6. Conidiogenous cells integrated in the distal ends of hyphae; conidia thick-walled, brown, smooth,<br />

-septate ......................................................................................................................................................................... Capnobotryella<br />

www.studiesinmycology.org<br />

11


Crous et al.<br />

6. Conidiophores short <strong>and</strong> frequently reduced to conidiogenous cells that proliferate percurrently via wide necks, giving rise to hyaline, 0(–2)-<br />

septate, broadly ellipsoidal conidia ................................................................................................................................................ Hortaea<br />

7. Conidia brown, with hyaline basal appendages; conidiomata pycnidial, conidiogenous cells phialidic, but also percurrent,<br />

subhyaline ....................................................................................................................................................................... Nothostrasseria<br />

7. Conidia brown, but basal appendages lacking, amero- to scolecospores ................................................................................................. 8<br />

8. Conidiomata pycnidial to acervular ............................................................................................................................................................ 9<br />

8. Conidiomata not enclosed by host tissue, fasciculate to sporodochial or solitary, hyphomycetous ........................................................ 10<br />

9. Conidia solitary, dry, without mucilaginous sheath .................................................................................................................... Readeriella<br />

9. Conidia catenulate, with persistant mucilaginous sheath ..................................................................................................... Staninwardia<br />

10. Conidiophores usually solitary, rarely densely fasciculate to synnematous (in vivo), penicillate, with a branched, apical conidiogenous<br />

apparatus giving rise to ramoconidia <strong>and</strong> branched chains of secondary conidia; scars not to slightly thickened <strong>and</strong> darkened-refractive<br />

..................................................................................................................................................................................................... Penidiella<br />

10. Conidiophores not penicillate, without a branched conidiogenous apparatus, in vivo fasciculate to sporodochial .................................. 11<br />

11. Biotrophic; fruiting composed of sporodochia <strong>and</strong> radiating layers of hyphae arising from the stromata, conidiophores arising from<br />

superficial sporodochia <strong>and</strong> radiating hyphae, conidiogenous cells unilocal, with conspicuous annellations, conidia solitary or in fragile<br />

disarticulating chains, aseptate or transversely 1–3-septate, usually with distinct frills, secession rhexolytic ............... Batcheloromyces<br />

. Biotrophic, leaf-inhabiting, with distinct, subepidermal to erumpent, well-developed sporodochia, or saxicolous, saprobic, sometimes<br />

causing opportunistic human infections; radiating layers of hyphae arising from sporodochia; conidiogenous cells without<br />

annellations; conidia in true simple or branched basipetal chains, transversely 1- to pluriseptate or with longitudinal <strong>and</strong> oblique septa<br />

(dictyosporous), occasionally distoseptate ...................................................................................................................... Catenulostroma<br />

To explain the arguments behind the selection <strong>and</strong> synonymies<br />

of some of these anamorphic genera, they are briefly discussed<br />

below:<br />

Acidomyces Baker et al., Appl. Environ. Microbiol. 70: 6270. 2004.<br />

(nom. inval.)<br />

Type species: Acidomyces richmondensis Baker et al., Appl.<br />

Environ. Microbiol. 70: 6270. 2004. (nom. inval.)<br />

Notes: <strong>The</strong> <strong>genus</strong> presently clusters among isolates in the<br />

Teratosphaeria clade based on sequences deposited in GenBank.<br />

Acidomyces lacks a Latin description <strong>and</strong> holotype specimen, <strong>and</strong><br />

is thus invalidly described. <strong>The</strong> <strong>genus</strong>, which was distinguished<br />

from other taxa based on its DNA phylogeny (Dothideomycetes),<br />

forms filamentous hyphae with disarticulating cells. It is unclear<br />

how it differs from Friedmanniomyces Onofri <strong>and</strong> Pseudotaeniolina<br />

J.L. Crane & Schokn.<br />

Batcheloromyces Marasas, P.S. van Wyk & Knox-Dav., J. S.<br />

African Bot. 41: 41. 1975.<br />

Type species: Batcheloromyces proteae Marasas, P.S. van Wyk &<br />

Knox-Dav., J. S. African Bot. 41: 43. 1975.<br />

Description: Crous et al. (2004a; figs 4–26).<br />

Notes: Batcheloromyces is presently circumscribed as a <strong>genus</strong><br />

that forms emergent hyphae, giving rise to superficial sporodochial<br />

plates, forming brown, verrucose, erect conidiophores that<br />

proliferate holoblastically, with ragged percurrent proliferations that<br />

become visible with age. Conidia are produced singly or in fragile,<br />

disarticulating chains, are brown, thick-walled, 0–3 transversely<br />

euseptate (though at times they appear as distoseptate). <strong>The</strong> <strong>genus</strong><br />

Batcheloromyces has in recent years been confused with Stigmina<br />

(Sutton & Pascoe 1989) on the basis that some collections showed<br />

conidiophores to give rise to solitary conidia only, though conidial<br />

catenulation was clearly illustrated by Taylor et al. (1999). In culture<br />

colonies tend to sporulate in a slimy mass (on OA), though a<br />

synanamorph can be seen (in B. leucadendri, Fig. 4) to sporulate<br />

via holoblastic conidiogenesis on hyphal tips of the aerial mycelium,<br />

forming elongate-globose to ellipsoid, muriformly septate, thickwalled<br />

conidia, that occur in clusters.<br />

<strong>The</strong> finding that Stigmina s. str. [based on S. platani (Fuckel)<br />

Sacc., the type species] is a generic synonym of Pseudocercospora<br />

Speg. (Crous et al. 2006a), <strong>and</strong> that the type species of<br />

Trimmatostroma (T. salicis, Fig. 5) belongs to the Helotiales<br />

(Fig. 1), raises the question of where to place stigmina- <strong>and</strong><br />

trimmatostroma-like anamorphs that reside in the Teratosphaeria<br />

clade. Although the stigmina-like species can be accommodated<br />

in Batcheloromyces (see Sutton & Pascoe 1989), a new <strong>genus</strong> is<br />

required for Teratosphaeria anamorphs that have a trimmatostromalike<br />

morphology. <strong>The</strong> recognition of Batcheloromyces <strong>and</strong> the<br />

introduction of a new anamorph <strong>genus</strong> for trimmatostroma-like<br />

anamorphs of Teratosphaeria are also morphologically justified.<br />

Batcheloromyces is easily distinguishable from Stigmina s.<br />

str. by its special structure of the fruiting body, composed of<br />

sporodochia <strong>and</strong> radiating layers of hyphae arising from the<br />

sporodochia <strong>and</strong> the conidia often formed in delicate disarticulating<br />

chains. Trimmatostroma-like anamorphs of Teratosphaeria are<br />

morphologically also sufficently distinct from Trimmatostroma s. str.<br />

(see notes under Catenulostroma Crous & U. Braun) as well as<br />

Batcheloromyces (see key above).<br />

Batcheloromyces eucalypti (Alcorn) Crous & U. Braun, comb.<br />

nov. MycoBank MB504503.<br />

Basionym: Stigmina eucalypti Alcorn, Trans. Brit. Mycol. Soc. 60:<br />

151. 1973.<br />

Capnobotryella Sugiy., in Sugiyama, Pleomorphic Fungi: <strong>The</strong><br />

Diversity <strong>and</strong> its Taxonomic Implications (Tokyo): 148. 1987.<br />

Type species: Capnobotryella renispora Sugiy., in Sugiyama,<br />

Pleomorphic Fungi: <strong>The</strong> Diversity <strong>and</strong> its Taxonomic Implications<br />

(Tokyo): 148. 1987.<br />

Description: Sugiyama & Amano (1987, figs 7.5–7.8).<br />

12


Phylogenetic lineages in the Capnodiales<br />

Fig. 4. Batcheloromyces leucadendri in vitro. A–B. Batcheloromyces state with synanamorph (arrows). C–D. Conidia occurring solitary or in short chains. Scale bar = 10 µm.<br />

Fig. 5. Trimmatostroma salicis. A. Sporodochia on twig. B–E. Chains of disarticulating conidia. Scale bars = 10 µm.<br />

Notes: <strong>The</strong> <strong>genus</strong> forms brown, septate, thick-walled hyphae, with<br />

ellipsoidal, 0–1-septate conidia forming directly on the hyphae, via<br />

minute phialides. Hambleton et al. (2003) also noted the occurrence<br />

of endoconidiation.<br />

Catenulostroma Crous & U. Braun, gen. nov. MycoBank<br />

MB504474.<br />

Etymology: Named after its catenulate conidia, <strong>and</strong> stromata giving<br />

rise to sporodochia.<br />

Hyphomycetes. Differt a Trimmatostromate habitu phytoparasitico, maculis<br />

formantibus, conidiophoris saepe fasciculatis, per stoma emergentibus vel habitu<br />

saxiphilo-saprophytico, interdum sejunctis ex mycosibus humanis.<br />

www.studiesinmycology.org<br />

13


Crous et al.<br />

Habit plant pathogenic, leaf-spotting or saxicolous-saprobic,<br />

occasionally isolated from opportunistic human mycoses.<br />

Mycelium internal <strong>and</strong> external; hyphae dark brown, septate,<br />

branched. Conidiomata in vivo vary from acervuli to sporodochia or<br />

fascicles of conidiophores arising from well-developed or reduced,<br />

pseudoparenchymatal stromata. Setae <strong>and</strong> hyphopodia absent.<br />

Conidiophores arising from hyphae or stromata, solitary, fasciculate<br />

to sporodochial, in biotrophic, plant pathogenic species emerging<br />

through stomata, little differentiated, semimacronematous, branched<br />

or not, continuous to septate, brown, smooth to verruculose.<br />

Conidiogenous cells integrated, terminal or conidiophores reduced<br />

to conidiogenous cells, holoblastic-thalloblastic, meristematic,<br />

unilocal, delimitation of conidium by a single septum with<br />

retrogressive delimitation of next conidium giving an unconnected<br />

chain of conidia, brown, smooth to verruculose, conidiogenous<br />

scars (conidiogenous loci) inconspicuous, truncate, neither<br />

thickened nor darkened. Conidia solitary or usually forming simple<br />

to branched basipetal chains of transversely to muriformly eu- or<br />

distoseptate, 1– to multiseptate, brown, smooth, verruculose to<br />

verrucose conidia, conidial secession schizolytic.<br />

Type species: Catenulostroma protearum (Crous & M.E. Palm)<br />

Crous & U. Braun, comb. nov.<br />

Description: Crous & Palm (1999), Crous et al. (2004a; figs 364–<br />

365).<br />

Notes: Catenulostroma contains several plant pathogenic species<br />

previously placed in Trimmatostroma, a morphologically <strong>similar</strong> but,<br />

based on its type species, phylogenetically distinct <strong>genus</strong> belonging<br />

to Helotiales (Fig. 1). Trimmatostroma s. str. is well-distinguished<br />

from most Catenulostroma species by being saprobic, living on<br />

twigs <strong>and</strong> branches of woody plants, or occasionally isolated<br />

from leaf litter, i.e., they are not associated with leaf spots. <strong>The</strong><br />

conidiomata of Trimmatostroma species are subepidermal,<br />

acervular-sporodochial with a well defined wall of textura angularis,<br />

little differentiated, micronematous conidiophores giving rise to long<br />

chains of conidia that disarticulate at the surface to form a greyblack<br />

to brown powdery mass. <strong>The</strong> generic affinity of other species<br />

assigned to Trimmatostroma, e.g. those having a lichenicolous<br />

habit, is unresolved.<br />

Trimmatostroma abietis Butin & Pehl (Butin et al. 1996) clusters<br />

together with the plant pathogenic Catenulostroma species, but<br />

differs from these species in having a more complex ecology.<br />

Trimmatostroma abietis is usually foliicolous on living or necrotic<br />

conifer needles on which characteristic acervuli to sporodochia<br />

with densely arranged, fasciculate fertile hyphae are formed,<br />

comparable to the fasciculate conidiomata of the plant pathogenic<br />

species of Catenulostroma (Butin et al. 1996: 205, fig. 1). Although<br />

not discussed by Butin et al. (1996), T. abietis needs to be compared<br />

to T. abietina Doherty, which was orginally described from Abies<br />

balsamea needles collected in Guelph, Canada (Doherty 1900).<br />

Morphologically the two species appear to be synonymous, except<br />

for reference to muriformly septate conidia, which is a feature not<br />

seen in vivo in the type of T. abietis. Furthermore, as this is clearly<br />

a species complex, this matter can only be resolved once fresh<br />

Canadian material has been collected to serve as epitype for T.<br />

abietina.<br />

Isolates from stone, agreeing with T. abietis in cultural,<br />

morphological <strong>and</strong> physiological characteristics, have frequently<br />

been found (Wollenzien et al. 1995, Butin et al. 1996, Gorbushina<br />

et al. 1996, Kogej et al. 2006, Krumbein et al. 1996). Furthermore,<br />

isolates from humans (ex skin lesions <strong>and</strong> ex chronic osteomycelitis<br />

of human patients) <strong>and</strong> Ilex leaves are known (Butin et al. 1996). De<br />

Hoog et al. (1999) included strains of T. abietis from stone, man <strong>and</strong><br />

Ilex leaves in molecular sequence analyses <strong>and</strong> demonstrated their<br />

genetical identity based on 5.8S rDNA <strong>and</strong> ITS2 data, but strains<br />

from conifer needles were not included. Furthermore, we consider<br />

T. abietis, as presently defined, to represent a species complex,<br />

with Dutch isolates from Pinus again appearing distinct from<br />

German Abies isolates, suggesting that different conifer genera<br />

could harbour different Catenulostroma species. Isolates from<br />

stone form stromatic, durable microcolonies, which are able to grow<br />

under extreme xerophilic environmental conditions. Cultural growth<br />

resembles that of other meristematic black yeasts (Butin et al. 1996,<br />

Kogej et al. 2006). Another fungus isolated from stone in Germany<br />

is in vitro morphologically close to C. abietis, but differs in forming<br />

conidia with oblique septa. Furthermore, a human pathogenic<br />

isolate from Africa clusters together with other Catenulostroma<br />

species. <strong>The</strong> habit <strong>and</strong> origin of this human pathogenic fungus in<br />

nature <strong>and</strong> its potential morphology on “natural” substrates, which<br />

typically deviates strongly from the growth in vitro, are still unknown.<br />

However, C. abietis, usually growing as a foliicolous <strong>and</strong> saxicolous<br />

fungus, has already shown the potential ability of Catenulostroma<br />

species to cause opportunistic human infections.<br />

Key to Catenulostroma species<br />

1. Conidia formed in basipetal chains, smooth, 4-celled, consisting of two basal cells with truncate lateral sides, each giving rise to a<br />

secondary globose apical cell, that can extend <strong>and</strong> develop additional septa, appearing as two lateral arms ................. C. excentricum<br />

1. Conidia variable in shape, but without two basal cells giving rise to two lateral arms ............................................................................... 2<br />

2. Conidia smooth or almost so, at most very faintly rough-walled; usually foliicolous on conifer needles or saxicolous, forming stromatic,<br />

xerophilic durable microcolonies on stone, occasionally causing opportunistic human infections ............................................................ 3<br />

2. Conidia distinctly verruculose to verrucose; plant pathogenic, forming leaf spots ..................................................................................... 5<br />

3. Conidia (8–)20–35(–60) × 4–5(–7) µm, 1–10-septate .................................................................................... C. chromoblastomycosum<br />

3. Conidia much shorter, 8–20 µm long, 0–5-septate .................................................................................................................................... 4<br />

4. Conidia 0–5 times transversely septate, mostly two-celled; usually foliicolous on conifer needles or saxicolous ....................... C. abietis<br />

4. Conidia 2–4 times transversely septate <strong>and</strong> often with 1–2 oblique septa; isolated from stone ........................................ C. germanicum<br />

5. Conidia rather broad, usually wider than 10 µm ........................................................................................................................................ 6<br />

14


Phylogenetic lineages in the Capnodiales<br />

Fig. 6. Catenulostroma chromoblastomycosum (type material). A. Sporodochium on pine needle in vitro. B–H. Chains of disarticulating conidia. Scale bars: A = 350, B, E, G,<br />

H = 10 µm.<br />

5. Conidia narrower, width below 10 µm ....................................................................................................................................................... 7<br />

6. Conidia distoseptate, rather long, (12–)25–35(–45) × (7–)10–15(–25) µm; conidiomata large, up to 250 µm diam, on Protea anceps<br />

............................................................................................................................................................................................... C. protearum<br />

6. Conidia euseptate, shorter, (9–)16–20(–36) × (10–)14–18(–27) µm; sporodochia 90–100 × 40–80 µm; on Protea gr<strong>and</strong>iceps<br />

................................................................................................................................................................................................. C. elginense<br />

7. Conidia 1- to multiseptate, (10–)15–17(–23) × (5–)6.5–7(–9) µm; on various Proteaceae .................................................. C. macowanii<br />

7. Conidia in vivo predominantly 1-septate, (8–)13–15(–21) × (3.5–)5.5–6(–8) µm; on Protea cynaroides<br />

......................................................................................................................................... C. microsporum (Teratosphaeria microspora)<br />

Catenulostroma abietis (Butin & Pehl) Crous & U. Braun, comb.<br />

nov. MycoBank MB504504.<br />

Basionym: Trimmatostroma abietis Butin & Pehl, Antonie van<br />

Leeuwenhoek 69: 204. 1996.<br />

Notes: Catenulostroma abietis needs to be compared to<br />

Trimmatostroma abietina Doherty (Abies balsamea needles<br />

Canada), which is either an older name for this species, or a closely<br />

related taxon. Presently T. abietina is not known from culture, <strong>and</strong><br />

needs to be recollected.<br />

Catenulostroma chromoblastomycosum Crous & U. Braun, sp.<br />

nov. MycoBank MB504505. Fig. 6.<br />

Etymology: Named after the disease symptoms observed due to<br />

opportunistic human infection.<br />

Differt a C. abieti et C. germanico conidiis longioribus, (8–)20–35(–60) × 4–5(–7)<br />

µm, 1–10-septatis.<br />

Description based on cultures sporulating on WA supplemented<br />

with sterile pine needles. Mycelium consisting of branched,<br />

septate, smooth to finely verruculose, medium to dark brown,<br />

thick-walled, 3–4 µm wide hyphae. Conidiomata brown, superficial,<br />

www.studiesinmycology.org<br />

15


Crous et al.<br />

Fig. 7. Catenulostroma germanicum (type material). A–D. Chains of disarticulating conidia in vitro. Scale bars = 10 µm.<br />

sporodochial, up to 350 µm diam. Conidiophores reduced to<br />

inconspicuous conidiogenous loci on hyphae, 2–4 µm wide, neither<br />

darkened nor thickened or refractive. Conidia occurring in branched<br />

chains, that tend to remain attached to each other, subcylindrical<br />

with subtruncate ends, straight to slightly curved, (8–)20–35(–60)<br />

× 4–5(–7) µm, 1–10-septate, medium brown, smooth to finely<br />

verruculose.<br />

Cultural characteristics: Colonies on PDA erumpent, spreading,<br />

slow growing, with sparse to moderate aerial mycelium <strong>and</strong> smooth,<br />

irregular, submerged margins; greenish black (surface).<br />

Specimen examined: Zaire, Pawa, isolated from man with chromoblastomycosis,<br />

Mar. 1997, V. de Brouwere, holotype <strong>CBS</strong> H-19935, culture ex-type <strong>CBS</strong> 597.97.<br />

Notes: Catenulostroma chromoblastomycosum was originally<br />

identified as an isolate of Stenella araguata Syd. <strong>The</strong> latter fungus<br />

is morphologically distinct, however, having much shorter <strong>and</strong><br />

narrower conidia, formed in acropetal chains, as well as quite<br />

different conidiogenous loci <strong>and</strong> conidial hila which are small,<br />

thickened <strong>and</strong> darkened.<br />

Catenulostroma elginense (Joanne E. Taylor & Crous) Crous &<br />

U. Braun, comb. nov. MycoBank MB504506.<br />

Basionym: Trimmatostroma elginense Joanne E. Taylor & Crous,<br />

Mycol. Res. 104: 633. 2000.<br />

Catenulostroma excentricum, see Teratosphaeria excentrica.<br />

Catenulostroma germanicum Crous & U. Braun, sp. nov.<br />

MycoBank MB504507. Fig. 7.<br />

Etymology: Named after the geographic location of its type strain<br />

in Germany.<br />

Differt a C. abieti conidiis 1–2 oblique septatis.<br />

Mycelium consisting of branched, septate, smooth, pale to<br />

medium brown, 2–4 µm wide hyphae, giving rise to conidial<br />

chains. Conidiophores integrated, subcylindrical, branched or not,<br />

septate, little differentiated, micronematous, 3–5 µm wide, 3- to<br />

multiseptate, medium brown, thick-walled; conidiogenous cells<br />

integrated, terminal, inconspicuous, unilocal, conidiogenous loci<br />

16


Phylogenetic lineages in the Capnodiales<br />

inconspicuous. Conidia in simple or branched basipetal chains,<br />

subcylindrical, straight to flexuous, (8–)10–15(–20) × 4–5(–6) µm,<br />

2–4 transversely septate or with 1–2 oblique septa, medium to dark<br />

brown, thick-walled, smooth.<br />

Cultural characteristics: Colonies on OA erumpent, spreading, with<br />

even, smooth margins <strong>and</strong> sparse to moderate aerial mycelium;<br />

olivaceous-grey, with iron-grey margins (surface). Colonies reaching<br />

12 mm diam after 1 mo at 25 °C in the dark; colonies fertile.<br />

Specimen examined: Germany (former West-Germany), isolated from stone, Oct.<br />

1988, J. Kuroczkin, holotype <strong>CBS</strong> H-19936, culture ex-type <strong>CBS</strong> 539.88.<br />

Notes: Catenulostroma germanicum was originally deposited as<br />

Taeniolina scripta (P. Karst.) P.M. Kirk. It is clearly distinct, however,<br />

as the latter fungus forms intricate, branched, brown conidia<br />

(Kirk 1981), unlike those of C. germanicum. Phylogenetically C.<br />

germanicum forms part of the C. abietis species complex.<br />

Catenulostroma macowanii (Sacc.) Crous & U. Braun, comb.<br />

nov. MycoBank MB504508.<br />

Basionym: Coniothecium macowanii Sacc., Syll. Fung. 4: 512.<br />

1886.<br />

≡ Coniothecium punctiforme G. Winter, Hedwigia 24: 33. 1885, non C.<br />

punctiforme Corda, Icones Fungorum (Prague) 1: 2. 1837.<br />

≡ Trimmatostroma macowanii (Sacc.) M.B. Ellis, More Dematiacous<br />

Hyphomycetes: 29. 1976.<br />

Catenulostroma microsporum, see Teratosphaeria microspora.<br />

Catenulostroma protearum (Crous & M.E. Palm) Crous & U.<br />

Braun, comb. nov. MycoBank MB504509.<br />

Basionym: Trimmatostroma protearum Crous & M.E. Palm, Mycol.<br />

Res. 103: 1303. 1999.<br />

Cibiessia Crous, Fungal Diversity 26: 151. 2007.<br />

Type species: Cibiessia dimorphospora Crous & C. Mohammed,<br />

Fungal Diversity 26: 151. 2007.<br />

Description: Crous et al. (2007b; figs 3–5).<br />

Notes: <strong>The</strong> <strong>genus</strong> Cibiessia was introduced to accommodate<br />

species with chains of disarticulating conidia (arthroconidia). Some<br />

species have been shown to have a Readeriella synanamorph.<br />

Devriesia Seifert & N.L. Nick., Can. J. Bot. 82: 919. 2004.<br />

Type species: Devriesia staurophora (W.B. Kendr.) Seifert & N.L.<br />

Nick., Canad. J. Bot. 82: 919. 2004.<br />

Description: Seifert et al. (2004; figs 2–42).<br />

Notes: <strong>The</strong> <strong>genus</strong> is characterised by producing chains of pale<br />

brown, subcylindrical to fusiform, 0–1-septate conidia with<br />

somewhat thickened, darkened hila, forming chlamydospores in<br />

culture, <strong>and</strong> being heat resistant. Morphologically they resemble<br />

taxa placed in Pseudocladosporium U. Braun (= Fusicladium<br />

Bonord.; Venturiaceae), though phylogenetically Devriesia is not<br />

allied to this family.<br />

Hortaea Nishim. & Miyaji, Jap. J. Med. Mycol. 26: 145. 1984.<br />

Type species: Hortaea werneckii (Horta) Nishim. & Miyaji, Jap. J.<br />

Med. Mycol. 26: 145. 1984.<br />

Description: de Hoog et al. (2000, illust. p. 721).<br />

www.studiesinmycology.org<br />

Notes: <strong>The</strong> <strong>genus</strong> forms brown, septate, thick-walled hyphae, with<br />

ellipsoidal, 0–1-septate (becoming muriformly septate), hyaline to<br />

pale brown conidia forming directly on the hyphae, via phialides<br />

with percurrent proliferation. Isolates of H. werneckii are restricted<br />

to tropical or subtropical areas, where they occur as halophilic<br />

saprobes, frequently being associated with tinea nigra of humans<br />

(de Hoog et al. 2000). <strong>The</strong> generic distinction with Capnobotryella<br />

is less clear, except that the latter tends to have darker, thick-walled<br />

conidia, <strong>and</strong> reduced, less prominent phialides.<br />

Penidiella Crous & U. Braun, gen. nov. MycoBank MB504463.<br />

Etymology: Named after its penicillate conidiophores.<br />

Differt a Periconiellae conidiophoris apice penicillato ex cellulis conidiogenis et<br />

ramoconidiis compositis, cellulis conidiogenis saepe 1–3(–4) locis conidiogenis,<br />

terminalibus vel subterminalibus, subdenticulatis, non vel subincrassatis, non vel<br />

leviter fuscatis-refractivis, ramoconidiis praesentibus, saepe numerosis, conidiis<br />

ramicatenatis.<br />

Mycelium consisting of branched, septate, smooth to verruculose,<br />

subhyaline to pale brown hyphae. Conidiophores macronematous,<br />

occasionally also with some micronematous conidiophores;<br />

macronematous conidiophores arising from superficial mycelium<br />

or stromata, solitary, fasciculate or in synnemata, erect, brown,<br />

thin- to thick-walled, smooth to finely verruculose; terminally<br />

penicillate, branched terminal part consisting of a conidiogenous<br />

apparatus composed of a series of conidiogenous cells <strong>and</strong>/or<br />

ramoconidia. Conidiogenous cells integrated, terminal, intercalary<br />

or pleurogenous, unbranched, pale to medium brown, smooth<br />

to finely verruculose, tapering to a flattened or rounded apical<br />

region or tips slightly inflated, polyblastic, sympodial, giving rise<br />

to a single or several sets of ramoconidia on different levels; with<br />

relatively few conidiogenous loci, 1–3(–4), terminal or subterminal,<br />

subdenticulate, denticle-like loci usually conical, terminally<br />

truncate, usually unthickened or at most very slightly thickened, not<br />

to slightly darkened or somewhat refractive. Conidia in branched<br />

acropetal chains. Ramoconidia 0–1-septate, pale to medium brown,<br />

smooth to verruculose, thin-walled, ellipsoidal, obovoid, fusiform,<br />

subcylindrical to obclavate; conidia subcylindrical, fusiform to<br />

ellipsoid-ovoid, 0–1-septate, pale olivaceous to brown, smooth to<br />

verruculose, thin-walled, catenate; hila truncate, unthickened or<br />

almost so, barely to somewhat darkened-refractive.<br />

Type species: Penidiella columbiana Crous & U. Braun, sp. nov.<br />

Notes: Three ramichloridium-like genera cluster within Capnodiales,<br />

namely Periconiella Sacc. [type: P. velutina (G. Winter) Sacc.],<br />

Ramichloridium Stahel ex de Hoog [type: R. apiculatum (J.H.<br />

Mill., Giddens & A.A. Foster) de Hoog] <strong>and</strong> Penidiella [type: P.<br />

columbiana Crous & U. Braun]. All three genera have brown,<br />

macronematous conidiophores with <strong>similar</strong> conidial scars. Within<br />

this complex, Ramichloridium is distinct in having a prominent<br />

rachis giving rise to solitary conidia. Periconiella <strong>and</strong> Penidiella<br />

are branched in the apical part of their conidiophores, <strong>and</strong> lack<br />

a rachis. In Periconiella conidia are solitary or formed in short,<br />

mostly simple chains, ramoconidia are lacking. <strong>The</strong> apical<br />

conidiogenous apparatus is composed of conidiogenous cells or<br />

branches with integrated, usually terminal conidiogenous cells,<br />

which are persistent. <strong>The</strong> conidiogenous cells are subcylindrical<br />

to somewhat clavate, usually not distinctly attenuated towards the<br />

tip, <strong>and</strong> have several, often numerous loci, aggregated or spread<br />

over the whole cell, terminal to usually lateral, flat, non-protuberant,<br />

not denticle-like, usually distinctly thickened <strong>and</strong> darkened, at least<br />

at the rim. In contrast, Penidiella has a quite distinct branching<br />

17


Crous et al.<br />

system, consisting of a single terminal conidiogenous cell giving<br />

rise to several ramoconidia that form secondary ramoconidia, etc.,<br />

or the branched apparatus is composed of several terminal <strong>and</strong><br />

sometimes lateral conidiogenous cells giving rise to sequences of<br />

ramoconidia (conidiogenous cells <strong>and</strong> ramoconidia are often barely<br />

distinguishable, with conidiogenous cells disarticulating, becoming<br />

ramoconidia). <strong>The</strong> branched apparatus may be loose to dense,<br />

metula-like. <strong>The</strong> conidiogenous cells have only few, usually 1–3<br />

(–4), terminal or subterminal subdenticulate loci, <strong>and</strong> ramoconidia<br />

are prominent <strong>and</strong> numerous, giving rise to branched chains of<br />

secondary conidia with flat-tipped hila. Some species of Penidiella<br />

with compact, metula-like branched apices are morphologically<br />

close to Metulocladosporiella Crous, Schroers, J.Z. Groenew., U.<br />

Braun & K. Schub. (Crous et al. 2006d). This <strong>genus</strong> encompasses<br />

two species of banana diseases belonging to Herpotrichiellaceae<br />

(Chaetothyriales), characterised by having conidiophore bases<br />

with rhizoid hyphal appendages <strong>and</strong> abundant micronematous<br />

conidiophores. Penidiella species with less pronounced penicillate<br />

apices, e.g. P. strumelloidea (Milko & Dunaev) Crous & U. Braun, are<br />

comparable with species of the <strong>genus</strong> Pleurotheciopsis B. Sutton<br />

(see Ellis 1976). <strong>The</strong> latter <strong>genus</strong> is distinct in having unbranched,<br />

often percurrently proliferating conidiophores, lacking ramoconidia<br />

<strong>and</strong> colourless conidia formed in simple chains.<br />

<strong>Cladosporium</strong> helicosporum R.F. Castañeda & W.B. Kendr.<br />

(Castañeda et al. 1997) is another penidiella-like fungus with<br />

terminally branched conidiophores, subdenticulate conidiogenous<br />

loci <strong>and</strong> conidia in long acropetal chains, but its affinity to Penidiella<br />

has still to be proven.<br />

Key to Penidiella species<br />

1. Conidiophores in vivo in well-developed, dense fascicles <strong>and</strong> distinct synnemata arising from a basal stroma; on fallen leaves of Ficus<br />

sp., Cuba ............................................................................................................................................................................... P. cubensis<br />

1. Conidiophores solitary, at most loosely aggregated .................................................................................................................................. 2<br />

2. Conidiophores with a terminal conidiogenous cell, often somewhat swollen, giving rise to several ramoconidia (on one level) that form<br />

chains of straight to distinctly curved conidia; isolated from leaf of Carex sp., Russia ..................................................... P. strumelloidea<br />

2. Penicillate apex of the conidiophores composed of a system of true branchlets, conidiogenous cells <strong>and</strong> ramoconidia or at least a sequence<br />

of ramoconidia on several levels; conidia usually straight ......................................................................................................................... 3<br />

3. Mycelium verruculose; long filiform conidiophores ending with a subdenticulate cell giving rise to sets of penicillate conidiogenous cells or<br />

ramoconidia which are barely distinguishable <strong>and</strong> turn into each other; ramoconidia <strong>and</strong> conidia consistently narrow, (1.5–)2(–2.5) µm<br />

wide, <strong>and</strong> aseptate, ramoconidia sometimes heterochromous; on living leaves of Nect<strong>and</strong>ra coriacea, Cuba ................... P. nect<strong>and</strong>rae<br />

3. Mycelium more or less smooth; penicillate apex at least partly with true branchlets; conidia wider, 2–5 µm, at least partly septate, uniformly<br />

pigmented ................................................................................................................................................................................................. 4<br />

4. Hyphae, conidiophores <strong>and</strong> conidia frequently distinctly constricted at the septa; penicillate apex of the conidiophores sparingly developed,<br />

branchlets more or less divergent; isolated from leaf litter of Smilax sp., Cuba .................................................................. P. rigidophora<br />

4. Hyphae <strong>and</strong> conidia without distinct constrictions at the septa; penicillate apex of the conidiophores usually well-developed, with abundant<br />

branchings ................................................................................................................................................................................................. 5<br />

5. Conidiophores short, up to 120 × 3–4 µm, frequently with intercalary conidiogenous cell, swollen at the conidiogenous portion just below<br />

the upper septum which render the conidiophores subnodulose to distinctly nodulose, apex ± loosely penicillate; conidia (4–)5–7(–8) µm<br />

long; occasionally with micronematous conidiophores; isolated from man with tinea nigra, Venezuela .......................... P. venezuelensis<br />

5. Conidiophores much longer, up to 800 µm, 7–9 µm wide at the base, not distinctly nodulose, penicillate apex loose to often more<br />

compact, tight, metula-like; conidia longer, 7–25 × 2–5 µm; micronematous conidiophores lacking; isolated from dead leaf of Paepalanthus<br />

columbianus, Colombia ........................................................................................................................................................ P. columbiana<br />

Penidiella columbiana Crous & U. Braun, sp. nov. MycoBank<br />

MB504510. Figs 8–9.<br />

Etymology: Named after its country of origin, Colombia.<br />

Mycelium ex hyphis ramosis, septatis, levibus, pallide brunneis, 2–3 µm latis<br />

compositum. Conidiophora ex hyphis superficialibus oriunda, penicillata, erecta,<br />

brunnea, crassitunicata, minute verruculosa, ad 800 µm longa, ad basim 7–9<br />

µm lata, ad apicem pluriramosa, ex ramibus diversibus et cellulis conidiogenis<br />

composita, ramibus primariis (–2) subcylindraceis, 1–7-septatis, 50–120 × 4–6 µm;<br />

ramibus secundariis (–2) subcylindraceis, 1–5-septatis, 40–60 × 4–6 µm; ramibus<br />

tertiariis et subsequentibus 1–4-septatis, 10–30 × 3–5 µm. Cellulae conidiogenae<br />

terminales vel laterales, non ramosae, 5–15 × 3–5 µm, modice brunneae, minute<br />

verruculosae, apicem versus attenuatae, truncatae vel rotundatae, polyblasticae,<br />

sympodiales, cicatrices conidiales incrassatae, sed leviter fuscatae et non<br />

refractivae. Ramoconidia 0–1-septata, modice brunnea, levia, ellipsoidea, obclavata<br />

vel obovoidea, cum 1–3 hilis terminalibus, 10–20 × 3–5 µm; conidia subcylindrica<br />

vel ellipsoidea, 0–1-septata, pallide brunnea, catenata (–10), hila truncata, non<br />

incrassata, vix vel leviter fuscata.<br />

Mycelium consisting of branched, septate, smooth, pale brown, 2–3<br />

µm wide hyphae. Conidiophores arising from superficial mycelium,<br />

terminally penicillate, erect, brown, wall up to 1 µm wide, almost<br />

smooth to finely verruculose, up to 800 µm tall, 7–9 µm wide at<br />

the base; conidiogenous region consisting of a series of branches<br />

composed of true branchlets, conidiogenous cells <strong>and</strong> ramoconidia,<br />

branched portion usually rather compact, even metula-like, but<br />

also looser, with divergent branches; primary branches (–2),<br />

subcylindrical, 1–7-septate, 50–120 × 4–6 µm; secondary branches<br />

(–2), subcylindrical, 1–5-septate, 40–60 × 4–6 µm; tertiary <strong>and</strong><br />

additional branches 1–4-septate, 10–30 × 3–5 µm. Conidiogenous<br />

cells terminal, intercalary or lateral, unbranched, 5–15 × 3–5<br />

µm, medium brown, finely verruculose, tapering to a flattened or<br />

rounded (frequently swollen) apical region, scars thickened, but<br />

only somewhat darkened, not refractive. Ramoconidia 0–1-septate,<br />

18


Phylogenetic lineages in the Capnodiales<br />

Fig. 8. Penidiella columbiana (type material). A. Conidiophores on pine needle in vitro. B–H. Conidiophores with chains of disarticulating conidia. Scale bars: A = 450, B–C =<br />

10 µm.<br />

medium brown, smooth, wall ≤ 1 µm wide, ellipsoidal to obclavate<br />

or obovoid, with 1–3 apical hila, 10–25 × 3–5 µm, ramoconidia<br />

with broadly truncate base, not or barely attenuated, up to 4 µm<br />

wide, or at least somewhat attenuated at the base, hila 1.5–3 µm<br />

wide. Conidia subcylindrical to ellipsoid, 0(–1)-septate, pale brown,<br />

in chains of up to 10, 7–15 × 2–3 µm, hila truncate, unthickened,<br />

barely to somewhat darkened, 1–2 µm wide.<br />

Cultural characteristics: Colonies on PDA erumpent, spreading,<br />

with moderate aerial mycelium <strong>and</strong> smooth, even, submerged<br />

margins; olivaceous-grey in central part, iron-grey in outer region<br />

(surface); colonies fertile.<br />

Specimen examined: Colombia, Páramo de Guasca, 3400 m alt., isolated from<br />

dead leaf of Paepalanthus columbianus (Eriocaulaceae), Aug. 1980, W. Gams,<br />

holotype <strong>CBS</strong> H-19937, culture ex-type <strong>CBS</strong> 486.80.<br />

Notes: This isolate was originally identified as belonging to the<br />

Stenella araguata species complex. <strong>The</strong> latter name has been<br />

somewhat confused in the literature, <strong>and</strong> has been incorrectly<br />

applied to isolates associated with opportunistic human infections<br />

(de Hoog et al. 2000). <strong>The</strong> “araguata” species complex is treated<br />

elsewhere in the volume (see Crous et al. 2007a – this volume).<br />

www.studiesinmycology.org<br />

Penidiella cubensis (R.F. Castañeda) U. Braun, Crous & R.F.<br />

Castañeda, comb. nov. MycoBank MB504511. Fig. 10.<br />

Basionym: <strong>Cladosporium</strong> cubense R.F. Castañeda, Fungi Cubenses<br />

II (La Habana): 4. 1987.<br />

In vivo: Colonies on fallen leaves, amphigenous, effuse, pilose,<br />

brown. Mycelium usually external, superficial, but also internal,<br />

composed of branched, septate, brown, thin-walled, smooth to<br />

rough-walled hyphae, 2–3 µm wide. Stromata present, 40–80 µm<br />

diam, brown, immersed. Conidiophores densely fasciculate or in<br />

distinct synnemata, arising from stromata, erect, synnemata up<br />

to about 1000 µm long <strong>and</strong> (10–)20–40(–50) µm wide, individual<br />

threads filiform, pluriseptate throughout, brown, thin-walled (≤<br />

0.5 µm), smooth or almost so to distinctly verruculose, apically<br />

penicillate. Conidiogenous cells integrated, terminal <strong>and</strong> intercalary,<br />

10–30 µm long, subcylindrical, terminal conidiogenous cells often<br />

slightly enlarged at the tip, with (1–)2–3(–4) terminal or subterminal<br />

subdenticulate conidiogenous loci, short conically truncate, 1–2<br />

µm diam, unthickened or almost so, but often slightly refractive or<br />

darkened-refractive, intercalary conidiogenous cells usually with a<br />

single lateral locus just below the upper septum, conidiogenous cells<br />

giving rise to a single set of primary ramoconidia, or a sequence<br />

of ramoconidia at different levels. Ramoconidia cylindrical to<br />

ellipsoid-fusoid, 8–18(–25) × 2–3 µm, aseptate, pale olivaceous,<br />

olivaceous-brown to brown, thin-walled, smooth or almost so to<br />

19


Crous et al.<br />

Fig. 9. Penidiella columbiana (type material). A.<br />

Conidiophores. B. Ramoconidia. C. Secondary conidia.<br />

Scale bar = 10 µm. U. Braun del.<br />

faintly verruculose, ramoconidia with broadly truncate base, barely<br />

narrowed, or ramoconidia more or less attenuated at the base, hila<br />

1–2 µm wide, unthickened or almost so, but often slightly refractive<br />

or darkened-refractive. Conidia in long acropetal chains, narrowly<br />

ellipsoid-ovoid, fusiform, 5–12(–15) × (1–)1.5–3 µm, aseptate, pale<br />

olivaceous to brownish, thin-walled, smooth to faintly rough-walled,<br />

ends attenuated, hila 1–1.5 µm wide, unthickened, not darkened,<br />

at most somewhat refractive.<br />

Specimen examined: Cuba, Guantánamo, Maisí, on fallen leaves of Ficus sp., 24<br />

Apr. 1986, M. Camino, holotype INIFAT C86/134 (HAL 2019 F, ex holotype).<br />

Notes: <strong>Cladosporium</strong> cubense was not available in culture <strong>and</strong><br />

molecular sequence data are not available, but type material could<br />

be re-examined <strong>and</strong> revealed that this species is quite distinct from<br />

<strong>Cladosporium</strong> s. str., but agreeing with the concept of the <strong>genus</strong><br />

Penidiella. Penidiella cubensis differs from all other species of this<br />

<strong>genus</strong> in having densely fasciculate conidiophores to synnematous<br />

conidiomata, arising from stromata.<br />

Penidiella nect<strong>and</strong>rae Crous, U. Braun & R.F. Castañeda, nom.<br />

nov. MycoBank MB504512. Fig. 11.<br />

Basionym: <strong>Cladosporium</strong> ferrugineum R.F. Castañeda, Fungi<br />

Cubenses II (La Habana): 4. 1987, homonym, non C. ferrugineum<br />

Allesch., 1895.<br />

In vivo: Colonies amphigenous, brown. Mycelium internal <strong>and</strong><br />

external, superficial, composed of sparingly branched hyphae,<br />

septate, 1–3 µm wide, pale olivaceous-brown or brown, thinwalled<br />

(≤ 0.5 µm), smooth or almost so to distinctly verruculose,<br />

fertile cells giving rise to conidiophores somewhat swollen at the<br />

branching point, up to 5 µm diam, <strong>and</strong> somewhat darker. Stromata<br />

lacking. Conidiophores erect, straight, filiform, up to 350 µm long,<br />

2.5–4 µm wide, pluriseptate throughout, brown, darker below <strong>and</strong><br />

paler above, thin-walled, smooth, apex penicillate, terminal cell of<br />

the conidiophore with 2–4 short denticle-like loci giving rise to sets<br />

of conidiogenous cells or ramoconidia that then form a sequence of<br />

new sets of ramoconidia on different levels, i.e., the loose to dense,<br />

metula-like branching system is composed of conidiogenous cells<br />

<strong>and</strong> ramoconidia which are often barely distinguishable <strong>and</strong> turn<br />

into each other; conidiogenous loci terminal or subterminal, usually<br />

1–3(–4), subdenticulate, 1–2 µm diam, conical, apically truncate,<br />

unthickened or almost so, not to somewhat darkened-refractive.<br />

Ramoconidia with truncate base, barely attenuated, or ramoconidia<br />

distinctly attenuated at the truncate base, up to 20 × 2 µm, aseptate,<br />

at the apex with 2–3(–4) subdenticulate hila, subcylindrical,<br />

20


Phylogenetic lineages in the Capnodiales<br />

very pale olivaceous, olivaceous-brown to brown, sometimes<br />

with different shades of brown (heterochromatous), thin-walled<br />

(≤ 0.5 µm), smooth to faintly verruculose. Conidia in long acropetal<br />

chains, narrowly ellipsoid-ovoid, fusiform to cylindrical, 5–16 ×<br />

(1.5–)2(–2.5) µm, aseptate, very pale olivaceous, olivaceousbrown<br />

to brown, thin-walled, smooth to very faintly rough-walled,<br />

primary conidia with rounded apex <strong>and</strong> trunacte base, somewhat<br />

attenuated, secondary conidia truncate at both ends, hila 1–1.5<br />

µm diam, unthickened or almost so, at most slightly darkenedrefractive.<br />

Cultural characteristics: Colonies on PDA slimy, smooth, spreading;<br />

aerial mycelium absent, margins smooth, irregular; surface black<br />

with patches of cream. Colonies reaching 20 mm diam after 1 mo<br />

at 25 °C in the dark; colonies sterile on PDA, SNA <strong>and</strong> OA.<br />

Specimen examined: Cuba, Matanzas, San Miguel de los Baños, isolated from<br />

living leaves of Nect<strong>and</strong>ra coriacea (Lauraceae), 24 Jan. 1987, R.F. Castañeda <strong>and</strong><br />

G. Arnold, holotype INIFAT C87/45, culture ex-type <strong>CBS</strong> 734.87, <strong>and</strong> HAL 2018 F<br />

(ex-holotype).<br />

Notes: Although the ex-type strain of <strong>Cladosporium</strong> ferrugineum<br />

is sterile, its LSU DNA phylogeny reveals it to be unrelated to<br />

<strong>Cladosporium</strong> s. str. (see Fig. 1 in Crous et al. 2007a – this volume).<br />

Based on a re-examination of the type material it could clearly be<br />

shown that the morphology of this species fully agrees with the<br />

concept of the new <strong>genus</strong> Penidiella, which is supported by its<br />

phylogenetic position within Capnodiales.<br />

Fig. 10. Penidiella cubensis (type material). A. Swollen stromatic base of synnema.<br />

B. Conidiophores. C. Ramoconidia. D. Secondary conidia. Scale bar = 10 µm.<br />

U. Braun del.<br />

Penidiella rigidophora Crous, R.F. Castañeda & U. Braun, sp.<br />

nov. MycoBank MB504513. Figs 12–13.<br />

≡ <strong>Cladosporium</strong> rigidophorum R.F. Castañeda, nom. nud. (herbarium<br />

name).<br />

Differt a specibus Penidiellae conidiophoris dimorphosis, hyphis et conidiis ad septa<br />

saepe distincte constrictis.<br />

Mycelium consisting of strongly branched, septate, smooth or<br />

almost so, pale olivaceous to medium brown, guttulate, commonly<br />

constricted at septa, 2–6 µm wide hyphae, swollen cells up to 8<br />

µm wide, wall up to 1(–1.5) µm wide. Conidiophores dimorphic.<br />

Macronematous conidiophores separate, erect, subcylindrical,<br />

predominantly straight to slightly curved, terminally loosely<br />

penicillate, up to 120 µm long, <strong>and</strong> 4–5 µm wide at the base, which<br />

is neither lobed nor swollen, <strong>and</strong> lacks rhizoids, up to 10-septate,<br />

medium to dark brown, wall up to 1(–1.5) µm wide. Micronematous<br />

conidiophores erect, subcylindrical, up to 40 µm tall, 3–4 µm wide,<br />

1–3-septate, pale to medium brown (concolorous with hyphae).<br />

Conidiogenous cells predominantly terminal, rarely intercalary,<br />

medium brown, smooth, subcylindrical, but frequently swollen at<br />

apex, 10–20 × 5–6 µm, loci (predominantly single in micronematous<br />

conidiophores, but up to 4 in macronematous conidiophores) flattipped,<br />

sub-denticulate or not, 1–1.5 µm wide, barely to slightly<br />

darkened <strong>and</strong> thickened-refractive. Conidia in branched chains,<br />

medium brown, verruculose, (appearing like small spines under<br />

light microscope), ellipsoid to cylindrical-oblong, up to 1(–1.5)<br />

µm wide, frequently constricted at septa, which turn dark with<br />

age; ramoconidia (10–)13–17(–25) × 3–4(–5) µm, 1(–3)-septate;<br />

secondary conidia (7–)8–10(–12) × 3–4(–5); hila unthickened to<br />

very slightly thickened <strong>and</strong> darkened, not refractive, (0.5–)1(–1.5)<br />

µm.<br />

Fig. 11. Penidiella nect<strong>and</strong>rae (type material). A. Conidiophores. B. Ramoconidia.<br />

C. Secondary conidia. Scale bar = 10 µm. U. Braun del.<br />

Cultural characteristics: Colonies on PDA erumpent, spreading,<br />

with lobate margins <strong>and</strong> moderate aerial mycelium; iron-grey<br />

(surface), with a greenish black margin; reverse greenish black.<br />

Colonies reaching 20 mm diam after 1 mo at 25 °C in the dark;<br />

colonies fertile.<br />

www.studiesinmycology.org<br />

21


Crous et al.<br />

Fig. 12. Penidiella rigidophora (type material). A–F. Micronematous conidiophores giving rise to chains of conidia. G–H. Macronematous conidiophores (note base in G, <strong>and</strong><br />

apex in H). I. Conidia. Scale bars = 10 µm.<br />

Specimen examined: Cuba, isolated from leaf litter of Smilax sp. (Smilacaceae), 6.<br />

Nov. 1994, R.F. Castañeda, holotype <strong>CBS</strong> H-19938, culture ex-type <strong>CBS</strong> 314.95.<br />

Notes: <strong>Cladosporium</strong> rigidophorum is a herbarium name, which was<br />

never validly published. <strong>The</strong> ex-type strain, however, represents a<br />

new species of Penidiella, for which a valid name with Latin diagnosis<br />

is herewith provided. This species is easily distinguishable from all<br />

other taxa of Penidiella by forming distinct constrictions at hyphal<br />

<strong>and</strong> conidial septa as well as micronematous conidiophores (except<br />

for P. venezuelensis in which a few micronematous conidiophores<br />

have been observed). It is also phylogenetically distinct from the<br />

other taxa of Penidiella (see Fig. 1 in Crous et al. 2007a – this<br />

volume).<br />

22


Phylogenetic lineages in the Capnodiales<br />

Penidiella strumelloidea (Milko & Dunaev) Crous & U. Braun,<br />

comb. nov. MycoBank MB504514. Figs 14–15.<br />

Basionym: <strong>Cladosporium</strong> strumelloideum Milko & Dunaev, Novosti<br />

Sist. Nizsh. Rast. 23: 134. 1986.<br />

Mycelium consisting of branched, septate, smooth, hyaline to<br />

pale olivaceous, 1–4 µm wide hyphae, sometimes constricted at<br />

somewhat darker septa. Conidiophores solitary, erect, arising from<br />

superficial mycelium, micronematous, i.e., reduced to conidiogenous<br />

cells, or macronematous, subcylindrical, straight to slightly curved,<br />

subcylindrical throughout or often somewhat attenuated towards the<br />

apex, 12–80 × (2–)2.5–4 µm, 0–6-septate, medium brown, smooth,<br />

wall ≤ 0.75 µm, penicillate apex formed by a terminal conidiogenous<br />

cell giving rise to a single set of ramoconidia. Conidiogenous cells<br />

terminal, integrated, subcylindrical, straight, 8–12 × 1.5–2(–2.5)<br />

µm, pale brown, thin-walled, smooth, apex obtusely rounded to<br />

somewhat clavate; loci terminal, occasionally subterminal or lateral,<br />

unthickened or almost so to slightly thickened <strong>and</strong> darkened, not<br />

refractive, 1–1.5 µm wide. Conidia in branched chains; ramoconidia<br />

subcylindrical, with 1–3 terminal loci, olivaceous-brown, smooth;<br />

secondary conidia ellipsoidal, with one side frequently straight <strong>and</strong><br />

the other convex, straight to slightly curved, (8–)10–12(–20) × 2(–<br />

3) µm, subhyaline to olivaceous-brown, smooth, thin-walled; hila<br />

unthickened or almost so to somewhat thickened <strong>and</strong> darkened,<br />

not refractive, 1 µm wide.<br />

Cultural characteristics: Colonies on PDA erumpent, spreading,<br />

with abundant, dense to woolly aerial mycelium, <strong>and</strong> uneven,<br />

feathery margins; surface pale olivaceous grey, reverse iron-grey.<br />

Colonies reaching 25 mm diam after 1 mo at 25 °C in the dark;<br />

colonies fertile.<br />

Fig. 13. Penidiella rigidophora (type material). A. Hyphae. B. Conidiophores. C.<br />

Ramoconidia. D. Secondary conidia. Scale bar = 10 µm. U. Braun del.<br />

Specimen examined: Russia, Yaroslavl Region, Rybinsk Reservoir, mouth of<br />

Sutka River, isolated from leaf of Carex sp. (Cyperaceae), from stagnant water, S.<br />

Ozerskaya, holotype BKMF-2534, culture ex-type <strong>CBS</strong> 114484.<br />

Fig. 14. Penidiella strumelloidea (type<br />

material). A–B. Micronematous conidiophores.<br />

C–D. Macronematous conidiophores. E–G.<br />

Conidia. Scale bars = 10 µm.<br />

www.studiesinmycology.org<br />

23


Crous et al.<br />

conidiophores, about 10–15 × 2–3 µm. Conidiogenous cells<br />

terminal <strong>and</strong> intercalary, unbranched, subcylindrical, 5–12 × 3–4<br />

µm, medium brown, smooth or almost so to finely verruculose,<br />

apex of conidiogenous cells frequently swollen, up to 6 µm diam,<br />

with 1–3(–4) flat-tipped, non to slightly thickened, non to slightly<br />

darkened-refractive loci, 1–1.5 µm wide, frequently appearing<br />

subdenticulate, up to 1.5 µm long, intercalary conidiogenous cells<br />

also slightly swollen at the conidiogenous portion just below the<br />

upper septum, which render the conidiophores subnodulose to<br />

nodulose, swellings round about the conidiophore axis or unilateral.<br />

Conidia ellipsoid-ovoid, subcylindrical, pale to medium olivaceousbrown<br />

or brown, finely verruculose, wall ≤ 0.5 µm wide, guttulate or<br />

not, occurring in branched chains. Ramoconidia 0–1(–3)-septate,<br />

5–15(–22) × 3–4(–5) µm, with 1–3 subdenticulate apical hila;<br />

secondary conidia 0(–1)-septate, ellipsoid, obovoid to irregular,<br />

(4–)5–7(–8) × (2–)2.5–3(–4) µm; hila non to slightly thickened, non<br />

to slightly darkened-refractive, (0.5–)1(–1.5) µm wide.<br />

Cultural characteristics: Colonies on OA erumpent, spreading,<br />

with dense, compact aerial mycelium, <strong>and</strong> even, smooth margins;<br />

olivaceous-grey (surface), margins iron-grey. Colonies reaching 22<br />

mm diam after 1 mo at 25 °C in the dark.<br />

Specimen examined: Venezuela, isolated from man with tinea nigra, Jan. 1975, D.<br />

Borelli, holotype <strong>CBS</strong> H-19934, culture ex-type <strong>CBS</strong> 106.75.<br />

Fig. 15. Penidiella strumelloidea (type material). A. Hyphae. B. Conidiophores. C.<br />

Ramoconidia. D. Secondary conidia. Scale bars = 10 µm. U. Braun del.<br />

Notes: Penidiella strumelloidea resembles other species of<br />

Penidiella by having penicillate conidiophores with a conidiogenous<br />

apparatus giving rise to branched conidial chains. It differs, however,<br />

from all other species of this <strong>genus</strong> in having a rather simple<br />

penicillate apex composed of a single terminal conidiogenous<br />

cell giving rise to one set of ramoconidia which form frequently<br />

somewhat curved conidia. It is also phylogenetically distinct from<br />

the other taxa of Penidiella (see Fig. 1 in Crous et al. 2007a – this<br />

volume).<br />

Penidiella venezuelensis Crous & U. Braun, sp. nov. MycoBank<br />

MB504515. Figs 16–17.<br />

Etymology: Named after the geographic location of its type strain,<br />

Venezuela.<br />

Differt a P. columbiana conidiophoris bevioribus et angustioribus, ad 120 × 3–4 µm,<br />

subnodulosis, apice plus minusve laxe penicillatis et conidiis brevioribus, (4–)5–7(–<br />

8) µm longis.<br />

Mycelium consisting of branched, septate, smooth to faintly<br />

rough-walled, thin-walled, subhyaline, pale olivaceous to medium<br />

brown, (1.5–)2–3 µm wide hyphae. Conidiophores solitary, erect,<br />

macronematous, subcylindrical, straight to flexuous to once<br />

geniculate, up to 120 µm long, 3–4 µm wide, 1–12-septate, pale to<br />

medium olivaceous-brown or brown, thin-walled (up to about 1 µm),<br />

terminally penicillate, branched portion composed of true branchlets<br />

<strong>and</strong>/or a single set or several sets of ramoconidia, branchlets up<br />

to 50 µm long; occasionally with a few additional micronematous<br />

Notes: <strong>The</strong> type culture of Penidiella venezuelensis was originally<br />

determined as Stenella araguata from which it is, however,<br />

quite distinct by having smooth mycelium, long penicillate<br />

conidiophores with subdenticulate conidiogenous loci, smaller<br />

conidia, <strong>and</strong> agreeing with the concept of the <strong>genus</strong> Penidiella. It is<br />

phylogenetically distinct from the other taxa of Penidiella (see Fig.<br />

1 in Crous et al. 2007a – this volume).<br />

Pseudotaeniolina J.L. Crane & Schokn., Mycologia 78: 88. 1986.<br />

? = Friedmanniomyces Onofri, Nova Hedwigia 68: 176. 1999.<br />

Type species: Pseudotaeniolina convolvuli (Esf<strong>and</strong>.) J.L. Crane &<br />

Schokn., Mycologia 78: 88. 1986.<br />

Description: Crane & Schoknecht (1986, figs 3–19).<br />

Notes: No cultures or sequence data are available of the type<br />

species, <strong>and</strong> Pseudotaeniolina globosa De Leo, Urzì & de Hoog was<br />

placed in Pseudotaeniolina based on its morphology <strong>and</strong> ecology.<br />

<strong>The</strong> <strong>genus</strong> Friedmanniomyces is presently known from two species<br />

(Selbmann et al. 2005). Morphologically Friedmanniomyces is<br />

<strong>similar</strong> to Pseudotaeniolina, but fresh material of Pseudotaeniolina<br />

convolvuli needs to be recollected before this can be clarified.<br />

Readeriella Syd. & P. Syd., Ann. Mycol. 6: 484. 1908.<br />

= Kirramyces J. Walker, B. Sutton & Pascoe, Mycol. Res. 96: 919. 1992.<br />

= Colletogloeopsis Crous & M.J. Wingf., Canad. J. Bot. 75: 668. 1997.<br />

Synanamorphs: Cibiessia Crous, Fungal Diversity 26: 151. 2007;<br />

also pseudocercospora-like, see Crous (1998).<br />

Type species: Readeriella mirabilis Syd. & P. Syd., Ann. Mycol. 6:<br />

484. 1908.<br />

Description: Crous et al. (2004b; figs 36–38).<br />

Notes: Several coelomycete genera are presently available<br />

to accommodate anamorphs of Capnodiales that reside in<br />

Teratosphaeriaceae, for which Readeriella is the oldest name. Other<br />

genera such as Phaeophleospora Rangel, Sonderhenia H.J. Swart<br />

& J. Walker <strong>and</strong> Lecanosticta Syd. belong to Mycosphaerellaceae.<br />

24


Phylogenetic lineages in the Capnodiales<br />

Fig. 16. Penidiella venezuelensis (type material). A. Microconidiophore. B. Apical part of macroconidiophore. C–F. Chains of conidia. Scale bars = 10 µm.<br />

Readeriella is polyphyletic within Teratosphaeriaceae. <strong>The</strong><br />

recognition <strong>and</strong> circumscription (synonymy) of this <strong>genus</strong> follows the<br />

principles for anamorph genera within Capnodiales as outlined in the<br />

introduction to this volume. <strong>The</strong> only unifying character is conidial<br />

pigmentation, <strong>and</strong> the mode of conidiogenesis. Conidiogenous<br />

cells range from mono- to polyphialides with periclinal thickening,<br />

to phialides with percurrent proliferation, as observed in the type<br />

species, R. mirabilis (Fig. 18). Within the form <strong>genus</strong> conidia vary<br />

from aseptate to multiseptate, smooth to rough, <strong>and</strong> have a range<br />

of synanamorphs. Readeriella mirabilis has a synanamorph with<br />

cylindrical, aseptate conidia, while other species of Readeriella<br />

again have Cibiessia synanamorphs (scytalidium-like, with chains<br />

of dry, disarticulating conidia), suggesting the conidial morphology<br />

to be quite plastic. A re-examination of R. readeriellophora Crous &<br />

Mansilla revealed pycnidia to form a central cushion on which the<br />

conidiogenous cells are arranged (Fig. 18). This unique feature is<br />

commonly known in genera such as Coniella Höhn. <strong>and</strong> Pilidiella<br />

Petr. & Syd. (Diaporthales) (Van Niekerk et al. 2004), <strong>and</strong> has never<br />

been observed among anamorphs of the Capnodiales. Another<br />

species of Readeriella, namely “Phaeophleospora” toledana Crous<br />

& Bills, again forms paraphyses interspersed among conidiogenous<br />

cells, a rare feature in this group of fungi, while several species<br />

Fig. 17. Penidiella venezuelensis (type material). A. Hypha. B. Micronematous<br />

conidiophores. C. Macronematous conidiophores. D–E. Ramoconidia. F. Secondary<br />

conidia. Scale bar = 10 µm. U. Braun del.<br />

www.studiesinmycology.org<br />

25


Crous et al.<br />

have conidiomata ranging from acervuli to pycnidia (Cortinas et<br />

al. 2006). Phylogenetically this coelomycete morphology, with<br />

its characteristic conidiogenesis, has evolved several times in<br />

Teratosphaeriaceae.<br />

Readeriella blakelyi (Crous & Summerell) Crous & U. Braun,<br />

comb. nov. MycoBank MB504516.<br />

Basionym: Colletogloeopsis blakelyi Crous & Summerell, Fungal<br />

Diversity 23: 342. 2006.<br />

Readeriella brunneotingens Crous & Summerell, sp. nov.<br />

MycoBank MB504517. Fig. 19.<br />

Etymology: Named after the diffuse brown pigment visible in agar<br />

when cultivated on MEA.<br />

Readeriellae gauchensi similis, sed coloniis viridi-atris et pigmento brunneo in agaro<br />

diffundente distinguenda.<br />

Leaf spots amphigenous, irregular specks up to 3 mm diam,<br />

medium brown with a thin, raised, concolorous border. Conidiomata<br />

amphigenous, substomatal, exuding conidia in black masses;<br />

conidiomata pycnidial in vivo <strong>and</strong> in vitro, globose, brown to black,<br />

up to 120 µm diam; wall consisting of 3–4 cell layers of brown<br />

cells of textura angularis. Conidiogenous cells brown, verruculose,<br />

aseptate, doliiform to ampulliform, or reduced to inconspicuous loci<br />

on hyphae (in vitro), proliferating percurrently near the apex, 5–7 ×<br />

3–5 µm; sympodial proliferation also observed in culture. Conidia<br />

brown, smooth to finely verruculose, ellipsoidal to subcylindrical,<br />

apex obtuse to subobtuse, tapering to a subtruncate or truncate<br />

base (1–1.5 µm wide) with inconspicuous, minute marginal frill,<br />

(5–)6–7(–8) × 2–3(–3.5) µm in vitro, becoming 1-septate; in older<br />

cultures becoming swollen, <strong>and</strong> up to 2-septate, 15 µm long <strong>and</strong><br />

5 µm wide.<br />

Cultural characteristics: Colonies on MEA reaching 20 mm diam<br />

after 2 mo at 25 °C; colonies erumpent, aerial mycelium sparse to<br />

absent, margins smooth but irregularly lobate; surface irregularly<br />

folded, greenish black, with profuse sporulation, visible as oozing<br />

black conidial masses; a diffuse dark-brown pigment is also<br />

produced, resulting in inoculated MEA plates appearing darkbrown.<br />

Specimen examined: Australia, Queensl<strong>and</strong>, Cairns, Eureka Creek, 48 km from<br />

Mareeba, S 17° 11’ 13.2”, E 145° 02’ 27.4”, 468 m, on leaves of Eucalyptus<br />

tereticornis, 26 Aug. 2006, P.W. Crous, <strong>CBS</strong>-H 19838 holotype, culture ex-type<br />

CPC 13303 = <strong>CBS</strong> 120747.<br />

Notes: Conidial dimensions of R. brunneotingens closely<br />

match those of Readeriella gauchensis (M.-N. Cortinas, Crous<br />

& M.J. Wingf.) Crous (Cortinas et al. 2006). <strong>The</strong> two species<br />

can be distinguished in culture, however, in that colonies of R.<br />

brunneotingens are greenish black in colour, sporulate profusely,<br />

<strong>and</strong> exude a diffuse, brown pigment into the agar, whereas colonies<br />

of R. gauchensis are more greenish olivaceous, <strong>and</strong> exude a yellow<br />

pigment into the agar (Cortinas et al. 2006).<br />

Readeriella considenianae (Crous & Summerell) Crous & U.<br />

Braun, comb. nov. MycoBank MB504518.<br />

Basionym: Colletogloeopsis considenianae Crous & Summerell,<br />

Fungal Diversity 23: 343. 2006.<br />

≡ Phaeophleospora destructans (M.J. Wingf. & Crous) Crous, F.A. Ferreira<br />

& B. Sutton, S. African J. Bot. 63: 113. 1997.<br />

Readeriella dimorpha (Crous & Carnegie) Crous & U. Braun,<br />

comb. nov. MycoBank MB504520.<br />

Basionym: Colletogloeopsis dimorpha Crous & Carnegie, Fungal<br />

Diversity 23: 345. 2006.<br />

Readeriella gauchensis (M.-N. Cortinas, Crous & M.J. Wingf.)<br />

Crous & U. Braun, comb. nov. MycoBank MB504521.<br />

Basionym: Colletogloeopsis gauchensis M.-N. Cortinas, Crous &<br />

M.J. Wingf., Stud. Mycol. 55: 143. 2006.<br />

Readeriella pulcherrima (Gadgil & M. Dick) Crous & U. Braun,<br />

comb. nov. MycoBank MB504522.<br />

Basionym: Septoria pulcherrima Gadgil & M. Dick, New Zeal<strong>and</strong> J.<br />

Bot. 21: 49. 1983.<br />

≡ Stagonospora pulcherrima (Gadgil & M. Dick) H.J. Swart, Trans. Brit.<br />

Mycol. Soc. 90: 285. 1988.<br />

= Cercospora eucalypti Cooke & Massee, Grevillea 18: 7. 1889.<br />

≡ Kirramyces eucalypti (Cooke & Massee) J. Walker, B. Sutton & Pascoe,<br />

Mycol. Res. 96: 920. 1992.<br />

≡ Phaeophleospora eucalypti (Cooke & Massee) Crous, F.A. Ferreira & B.<br />

Sutton, S. African J. Bot. 63: 113. 1997.<br />

Notes: <strong>The</strong> epithet “eucalypti” is preoccupied by Readeriella<br />

eucalypti (Gonz. Frag.) Crous (Summerell et al., 2006), <strong>and</strong> thus<br />

the synonym “pulcherrima” becomes the next available name for<br />

this species.<br />

Readeriella readeriellophora, see Teratosphaeria readeriellophora.<br />

Fig. 18.<br />

Readeriella stellenboschiana (Crous) Crous & U. Braun, comb.<br />

nov. MycoBank MB504523.<br />

Basionym: Colletogloeopsis stellenboschiana Crous, Stud. Mycol.<br />

55: 110. 2006.<br />

Readeriella zuluensis (M.J. Wingf., Crous & T.A. Cout.) Crous &<br />

U. Braun, comb. nov. MycoBank MB504524.<br />

Basionym: Coniothyrium zuluense M.J. Wingf., Crous & T.A. Cout.,<br />

Mycopathologia 136: 142. 1997.<br />

≡ Colletogloeopsis zuluensis (M.J. Wingf., Crous & T.A. Cout.) M.-N.<br />

Cortinas, M.J. Wingf. & Crous (zuluense), Mycol. Res. 110: 235. 2006.<br />

Staninwardia B. Sutton, Trans. Br. Mycol. Soc. 57: 540. 1971.<br />

Type species: Staninwardia breviuscula B. Sutton, Trans. Br. Mycol.<br />

Soc. 57: 540. 1971.<br />

Description: Sutton (1971; fig. 1).<br />

Notes: <strong>The</strong> <strong>genus</strong> Staninwardia presently contains two species,<br />

namely S. breviuscula <strong>and</strong> Staninwardia suttonii Crous & Summerell<br />

(Summerell et al. 2006), though its placement in Capnodiales was<br />

less well resolved. <strong>The</strong> <strong>genus</strong> forms acervuli on brown leaf spots,<br />

with brown, catenulate conidia covered in a mucilaginous sheath.<br />

Readeriella destructans (M.J. Wingf. & Crous) Crous & U. Braun,<br />

comb. nov. MycoBank MB504519.<br />

Basionym: Kirramyces destructans M.J. Wingf. & Crous, S. African<br />

J. Bot. 62: 325. 1996.<br />

26


Phylogenetic lineages in the Capnodiales<br />

Fig. 18. A–E. Readeriella mirabilis. A. Conidium with conidial cirrus. B. Conidiogenous cells with percurrent proliferation. C. Macroconidia. D. Slightly pigmented, verruculose<br />

conidiogenous cell. E. Macro- <strong>and</strong> microconidia. F–I. Readeriella readeriellophora (type material). F. Colony on OA. G. Central stromatal tissue giving rise to conidiophores. H.<br />

Conidiogenous cells. I. Conidia. Scale bars = 10 µm.<br />

Fig. 19. Readeriella brunneotingens (type material). A. Leaf spot. B. Colony on MEA. C–D. Conidia. Scale bar = 10 µm.<br />

www.studiesinmycology.org<br />

27


Crous et al.<br />

Schizothyriaceae clade<br />

Schizothyrium Desm., Ann. Sci. Nat., Bot., sér. 3: 11. 1849.<br />

Type species: Schizothyrium acerinum Desm., Ann. Sci. Nat., Bot.,<br />

sér. 3: 11. 1849.<br />

Description: Batzer et al. (2007; figs 3–7).<br />

Notes: Species of Schizothyrium (Schizothyriaceae) have<br />

Zygophiala E.W. Mason anamorphs, <strong>and</strong> were recently shown to<br />

be allied to Mycosphaerellaceae (Batzer et al. 2007). Although<br />

species of Schizothyrium have thyrothecia, they cluster among<br />

genera with pseudothecial ascomata, questioning the value of this<br />

character at the family level. Based on its bitunicate asci <strong>and</strong> 1-<br />

septate ascospores, the teleomorph is comparable to others in the<br />

Capnodiales.<br />

Mycosphaerellaceae clade<br />

Mycosphaerella subclade<br />

Mycosphaerella Johanson, Öfvers. Förh. Kongl. Svenska<br />

Vetensk.-Akad. 41(9): 163. 1884.<br />

Type species: Mycosphaerella punctiformis (Pers. : Fr.) Starbäck,<br />

Bih. Kongl. Svenska Vetensk.-Akad. H<strong>and</strong>l. 15(3, 2): 9. 1889.<br />

Anamorph: Ramularia endophylla Verkley & U. Braun, Mycol. Res.<br />

108: 1276. 2004.<br />

Description: Verkley et al. (2004; figs 3–16).<br />

Notes: <strong>The</strong> <strong>genus</strong> Mycosphaerella has in the past been linked to 23<br />

anamorph genera (Crous et al. 2000), while additional genera have<br />

been linked via DNA-based studies, bringing the total to at least 30<br />

genera (Crous & Braun 2003, Crous et al. 2007b). However, based<br />

on ITS <strong>and</strong> SSU DNA phylogenetic studies <strong>and</strong> a reassessment of<br />

morphological characters <strong>and</strong> conidiogenesis, several anamorph<br />

genera have recently been reduced to synonymy (Crous & Braun<br />

2003, Crous et al. 2006a). Furthermore, the DNA sequence data<br />

generated to date clearly illustrate that the anamorph genera in<br />

Mycosphaerella are polyphyletic, residing in several clades within<br />

Mycosphaerella. If future collections not known from culture or DNA<br />

sequences are to be described in form genera, we recommend that<br />

the concepts as explained in Crous & Braun (2003) be used until<br />

such stage as they can be placed in Mycosphaerella, pending<br />

a modification of Art. 59 of the International Code of Botanical<br />

Nomenclature. <strong>The</strong> <strong>genus</strong> Mycosphaerella <strong>and</strong> its anamorphs<br />

represent a future topical issue of the Studies in Mycology, <strong>and</strong> will<br />

thus be treated separately.<br />

Dissoconium subclade<br />

Dissoconium de Hoog, Oorschot & Hijwegen, Proc. K. Ned. Akad.<br />

Wet., Ser. C, Biol. Med. Sci. 86(2): 198. 1983.<br />

Type species: Dissoconium aciculare de Hoog, Oorschot &<br />

Hijwegen, Proc. K. Ned. Akad. Wet., Ser. C, Biol. Med. Sci. 86(2):<br />

198. 1983.<br />

? = Uwebraunia Crous & M.J. Wingf., Mycologia 88: 446. 1996.<br />

Teleomorph: Mycosphaerella-like.<br />

Description: de Hoog et al. (1983), Crous (1998), Crous et al.<br />

(2004b; figs 3–10).<br />

Notes: <strong>The</strong> <strong>genus</strong> Dissoconium presently encompasses six<br />

species (Crous et al. 2007b), of which two, M. lateralis Crous &<br />

M.J. Wingf. (D. dekkeri de Hoog & Hijwegen), <strong>and</strong> M. communis<br />

Crous & Mansilla (D. commune Crous & Mansilla) are also known<br />

from their Mycosphaerella-like teleomorphs. No teleomorph <strong>genus</strong><br />

will be introduced for this clade, however, until more sexual species<br />

have been collected to help clarify the morphological features of<br />

this <strong>genus</strong>. A further complication lies in the fact that yet other<br />

species, morphologically distinct from Dissoconium, also cluster in<br />

this clade (Crous, unpubl. data).<br />

“Passalora” zambiae subclade<br />

“Passalora” zambiae Crous & T.A. Cout., Stud. Mycol. 50: 209.<br />

2004.<br />

Description: Crous et al. (2004b; figs 32–33).<br />

Notes: This fungus was placed in the form <strong>genus</strong> “Passalora”<br />

based on its smooth mycelium, giving rise to conidiophores forming<br />

branched chains of brown conidia with thickened, darkened,<br />

refractive hila. Although derived from single ascospores, the<br />

teleomorph material was lost, <strong>and</strong> thus it needs to be recollected<br />

before the relavance of its phylogenetic position can be fully<br />

understood.<br />

Additional teleomorph genera considered<br />

Coccodinium A. Massal., Atti Inst. Veneto Sci. Lett. Arti, Série 2,<br />

5: 336. 1860. (Fig. 20).<br />

Type species: Coccodinium bartschii A. Massal., Atti Inst. Veneto<br />

Sci. Lett. Arti, Série 2, 5: 337. 1860.<br />

Description: Eriksson (1981, figs 34–35).<br />

Notes: <strong>The</strong> <strong>genus</strong> Coccodinium (Coccodiniaceae) is characterised<br />

by having ascomata that are sessile on a subiculum, or somewhat<br />

immersed, semiglobose, collapsed when dry, brownish, uniloculate,<br />

with a centrum that stains blue in IKI (iodine potassium iodide).<br />

Asci are bitunicate, stalked, 8-spored, saccate, <strong>and</strong> have a thick,<br />

undifferentiated endotunica. Periphyses <strong>and</strong> periphysoids are<br />

well-developed <strong>and</strong> numerous. Ascospores are elongate, fusiform,<br />

ellipsoidal or clavate, transversely septate or muriform, hyaline or<br />

brownish (Eriksson 1981), <strong>and</strong> lack a mucous sheath. Based on a<br />

SSU sequence (GenBank accession U77668) derived from a strain<br />

identified as C. bartschii (Winka et al. 1998), Coccodinium appears<br />

to be allied to the taxa treated here in Teratosphaeria. Freshly<br />

collected cultures are relatively slow growing, <strong>and</strong> on MEA they<br />

form erumpent round, black colonies with sparse hyphal growth. On<br />

the surface of these colonies hyphal str<strong>and</strong>s, consisting of brown,<br />

globose cells, give rise to conidia. Older cells (up to 15 µm diam)<br />

become fertile, giving rise to 1–3 conidia via inconspicuous phialidic<br />

loci. Conidia are fusoid-ellipsoidal to clavate, 3–5-septate, becoming<br />

constricted at the transverse septa, apex obtuse, base subtruncate,<br />

guttulate, smooth, widest in the upper third of the conidium, 15–<br />

40 × 4–7 µm. Phylogenetically Coccodinium is thus allied to the<br />

Chaetothyriales (Fig. 1), <strong>and</strong> not the Teratosphaeriaceae.<br />

28


Phylogenetic lineages in the Capnodiales<br />

Fig. 20. Coccodinium bartschii. A. Ascomata on host. B. Ostiolar area. C. Periphysoids. D–E. Ascospores shot onto agar. F–I. Asci with thick ectotunica. J–K. Young ascospores.<br />

L–M. Mature ascospores. N. Colony on MEA. O–Q. Conidiogenous cells giving rise to conidia. R–S. Conidia. Scale bars: A, N = 250, B, D, F–G, I, L–M, O = 10 µm.<br />

www.studiesinmycology.org<br />

29


Crous et al.<br />

Fig. 21. Stigmidium schaereri. A. Lichenicolous habit on Dacampia hookeri. B. Vertical section through an ascoma. C–D. Asci. E–G. Ascospores. H. Older, brown ascospores.<br />

Scale bars = 10 µm.<br />

Stigmidium Trevis., Consp. Verruc.: 17. 1860. (Fig. 21).<br />

Type species: Stigmidium schaereri (A. Massal.) Trevis., Consp.<br />

Verruc.: 17. 1860.<br />

Description: Roux & Triebel (1994, figs 47–50).<br />

Notes: <strong>The</strong> type species of the <strong>genus</strong> is lichenicolous, characterised<br />

by semi-immersed, black, globose ascomata with ostiolar<br />

periphyses <strong>and</strong> periphysoids. Asci are 8-spored, fasciculate,<br />

bitunicate, (endotunica not giving a special reaction in Congo red<br />

or toluidine blue). Ascospores are fusoid-ellipsoidal, medianly 1-<br />

septate, guttulate, thin-walled, lacking a sheath. Presently no<br />

culture is available, <strong>and</strong> thus the placement of Stigmidium remains<br />

unresolved.<br />

Discussion<br />

From the LSU sequence data presented here, it is clear that<br />

Mycosphaerella is not monophyletic as previously suggested (Crous<br />

et al. 2001, Goodwin et al. 2001). <strong>The</strong> first step to circumscribe<br />

natural genera within this complex was taken by Braun et al. (2003),<br />

who separated <strong>Cladosporium</strong> anamorphs from this complex,<br />

<strong>and</strong> erected Davidiella (Davidiellaceae; Schoch et al. 2006) to<br />

accommodate their teleomorphs. <strong>The</strong> present study reinstates<br />

the <strong>genus</strong> Teratosphaeria for a clade of largely extremotolerant<br />

fungi (Selbmann et al. 2005) <strong>and</strong> foliar pathogens of Myrtaceae<br />

<strong>and</strong> Proteaceae (Crous et al. 2004a, b, 2006b, 2007b), <strong>and</strong> further<br />

separates generic subclades within the Mycosphaerellaceae,<br />

while Batzer et al. (2007) again revealed Schizothyrium Desm.<br />

(Schizothyriaceae) to cluster within the Mycosphaerellaceae. Our<br />

results, however, provide support for recognition of Schizothyrium as<br />

a distinct <strong>genus</strong>, although Schizothyriaceae was less well supported<br />

as being separate from Mycosphaerellaceae (Capnodiales).<br />

Although pleomorphism represents a rather unstudied<br />

phenomenon in this group of fungi, it has been observed in several<br />

species. Within the Teratosphaeria clade, Crous et al. (2007b)<br />

recently demonstrated teleomorphs to have Readeriella <strong>and</strong><br />

Cibiessia synanamorphs, while the black yeast genera that belong<br />

to this clade, commonly have more than one anamorph state in<br />

culture. <strong>The</strong> present study also revealed Readeriella mirabilis to<br />

have two conidial types in culture, <strong>and</strong> to be highly plastic regarding<br />

its mode of conidiogenesis, <strong>and</strong> Readeriella to be the oldest generic<br />

name available for a large group of leaf-spotting coelomycetes in<br />

the Teratosphaeriaceae (Capnodiales).<br />

Although not commonly documented, there are ample<br />

examples of synanamorphs in Capnodiales. Within Mycosphaerella,<br />

Beilharz et al. (2004) described Passalora perplexa Beilharz,<br />

Pascoe, M.J. Wingf. & Crous as a species with a coelomycete<br />

<strong>and</strong> yeast synanamorph, while Crous & Corlett (1998) described<br />

Mycosphaerella stigmina-platani F.A. Wolf to have a Cercostigmina<br />

U. Braun <strong>and</strong> Xenostigmina Crous synanamorph, <strong>and</strong> recent<br />

collections also revealed the presence of a <strong>similar</strong> species that has<br />

typical “Stigmina” (distoseptate conidia) <strong>and</strong> Pseudocercospora<br />

(euseptate conidia) synanamorphs (Crous, unpubl. data), <strong>and</strong><br />

Crous (1998) reported Readeriella epicoccoides (coelomycete) to<br />

30


Phylogenetic lineages in the Capnodiales<br />

have a Cercostigmina (hyphomycete) synanamorph in culture.<br />

Although the Mycosphaerella complex encompasses<br />

thous<strong>and</strong>s of names, it may appear strange that it is only now that<br />

more clarity is obtained regarding the phylogenetic relationships<br />

among taxa in this group. This is partly due to the fact that these<br />

organisms are cultivated with difficulty, <strong>and</strong> also that the first paper<br />

to address the taxonomy of this complex based on DNA sequence<br />

data was only relatively recently published (Stewart et al. 1999).<br />

In the latter study, the <strong>genus</strong> Paracercospora Deighton (scars<br />

minutely thickened along the rim), was shown to be synonymous<br />

with the older <strong>genus</strong> Pseudocercospora. Similarily, Crous et al.<br />

(2001) showed that Cercostigmina (rough, irregular percurrent<br />

proliferations) was also synonymous with Pseudocercospora.<br />

This led Crous & Braun (2003) to conclude that conidiomatal type,<br />

conidial catenulation, septation <strong>and</strong> proliferation of conidiogenous<br />

cells were of less importance in separating species at the generic<br />

level. Mycovellosiella Rangel <strong>and</strong> Phaeoramularia Munt.-Cvetk.<br />

were subsequently reduced to synonymy with the older name,<br />

Passalora Fr., <strong>and</strong> characters identified as significant at the generic<br />

level were pigmentation (Cercospora vs. Passalora), scar structure<br />

(Passalora vs. Pseudocercospora), <strong>and</strong> verruculose superficial<br />

hyphae (Stenella vs. Passalora). Due to the unavailability of<br />

cultures, no decision was made regarding Stenella (verrucose<br />

conidia <strong>and</strong> mycelium), Stigmina (distoseptate conidia), <strong>and</strong><br />

several other, less well-known genera such as Asperisporium<br />

Maubl., Denticularia Deighton, Distocercospora N. Pons & B.<br />

Sutton, Prathigada Subram., Ramulispora, Pseudocercosporidium<br />

Deighton, Stenellopsis B. Huguenin <strong>and</strong> Verrucisporota D.E. Shaw<br />

& Alcorn. In a recent study, however, Crous et al. (2006a) were<br />

able to show that Phaeoisariopsis (synnemata, conidia with slightly<br />

thickened hila) <strong>and</strong> Stigmina (distoseptate conidia) were also<br />

synonyms of Pseudocercospora.<br />

<strong>The</strong> present study shows that most anamorph genera are<br />

polyphyletic within Teratosphaeria, <strong>and</strong> paraphyletic within<br />

Capnodiales. In some cases, generic concepts of anamorphs<br />

based on morphology <strong>and</strong> conidium ontogeny conform well with<br />

phylogenetic relationships, though this is not true in all cases due<br />

to convergence. Nevertheless, anamorphs still convey valuable<br />

morphological information that is contained in the anamorph name,<br />

<strong>and</strong> naming anamorphs continue to provide a practical system to<br />

identify the various asexual taxa encountered.<br />

Acknowledgements<br />

We gratefully acknowledge several colleagues in different countries who have<br />

collected the material studied here, without which this work would not have been<br />

possible. We thank M. Vermaas for preparing the photographic plates. M. Grube is<br />

gratefully acknowledged for sending us fresh material of Stigmidium to include in<br />

this study, <strong>and</strong> D. Triebel <strong>and</strong> M. Grube are thanked for advice regarding the strain of<br />

hamathecial tissue in Stigmidium. K.A. Seifert is thanked for numerous discussions<br />

about anamorph concepts <strong>and</strong> advice pertaining to this paper, <strong>and</strong> for recollecting<br />

Coccodinium bartschii to enable us to clarify its phylogeny. J.K. Stone is thanked for<br />

commenting on an earlier draft version of this paper.<br />

References<br />

Aptroot A (2006). Mycosphaerella <strong>and</strong> its anamorphs: 2. Conspectus of<br />

Mycosphaerella. <strong>CBS</strong> Biodiversity Series 5: 1–231.<br />

Barnes I, Crous PW, Wingfield BD, Wingfield MJ (2004). Multigene phylogenies<br />

reveal that red b<strong>and</strong> needle blight of Pinus is caused by two distinct species of<br />

Dothistroma, D. septosporum <strong>and</strong> D. pini. Studies in Mycology 50: 551–565.<br />

Barr ME (1972). Preliminary studies on the Dothideales in temperate North America.<br />

Contributions from the University of Michigan Herbarium 9: 523–638.<br />

Batzer JC, Mercedes Diaz Arias M, Harrington TC, Gleason ML, Groenewald JZ,<br />

Crous PW (2007). Four species of Zygophiala (Schizothyriaceae, Capnodiales)<br />

are associated with the sooty blotch <strong>and</strong> flyspeck complex on apple. Mycologia:<br />

in press.<br />

Beilharz VC, Pascoe IG, Wingfield MJ, Tjahjono B, Crous PW (2004). Passalora<br />

perplexa, an important pleoanamorphic leaf blight pathogen of Acacia<br />

crassicarpa in Australia <strong>and</strong> Indonesia. Studies in Mycology 50: 471–479.<br />

Braun U, Crous PW, Dugan F, Groenewald JZ, Hoog GS de (2003). Phylogeny <strong>and</strong><br />

taxonomy of cladosporium-like hyphomycetes, including Davidiella gen. nov.,<br />

the teleomorph of <strong>Cladosporium</strong> s.str. Mycological Progress 2: 3–18.<br />

Butin H, Pehl L, Hoog GS de, Wollenzien U (1996). Trimmatostroma abietis sp. nov.<br />

(hyphomycetes) <strong>and</strong> related species. Antonie van Leeuwenhoek 69: 203–209.<br />

Castañeda RF, Kendrick B, Gené J (1997). Notes on conidial fungi. XIII. A new<br />

species of <strong>Cladosporium</strong> from Cuba. Mycotaxon 63: 183–187.<br />

Chupp C (1954). A monograph of the fungus <strong>genus</strong> Cercospora. Ithaca, New York.<br />

Published by the author.<br />

Cortinas M-N, Crous PW, Wingfield BD, Wingfield MJ (2006). Multi-gene phylogenies<br />

<strong>and</strong> phenotypic characters distinguish two species within the Colletogloeopsis<br />

zuluensis complex associated with Eucalyptus stem cankers. Studies in<br />

Mycology 55: 133–146.<br />

Crane JL, Schoknecht JD (1986). Revision of Torula <strong>and</strong> Hormiscium species.<br />

New names for Hormiscium undulatum, Torula equina, <strong>and</strong> Torula convolvuli.<br />

Mycologia 78: 86–91.<br />

Crous PW (1998). Mycosphaerella spp. <strong>and</strong> their anamorphs associated with leaf<br />

spot diseases of Eucalyptus. Mycologia Memoir 21: 1–170.<br />

Crous PW, Aptroot A, Kang JC, Braun U, Wingfield MJ (2000). <strong>The</strong> <strong>genus</strong><br />

Mycosphaerella <strong>and</strong> its anamorphs. Studies in Mycology 45: 107–121.<br />

Crous PW, Braun U (2003). Mycosphaerella <strong>and</strong> its anamorphs. 1. Names published<br />

in Cercospora <strong>and</strong> Passalora. <strong>CBS</strong> Biodiversity Series 1: 1–571.<br />

Crous PW, Braun U, Schubert K, Groenewald JZ (2007a). Delimiting <strong>Cladosporium</strong><br />

from morphologically <strong>similar</strong> genera. Studies in Mycology 58: 33–56.<br />

Crous PW, Corlett M (1998). Reassessment of Mycosphaerella spp. <strong>and</strong> their<br />

anamorphs occurring on Platanus. Canadian Journal of Botany 76: 1523–<br />

1532.<br />

Crous PW, Denman S, Taylor JE, Swart L, Palm ME (2004a). Cultivation <strong>and</strong><br />

diseases of Proteaceae: Leucadendron, Leucospermum <strong>and</strong> Protea. <strong>CBS</strong><br />

Biodiversity Series 2: 1–228.<br />

Crous PW, Groenewald JZ (2006a). Mycosphaerella alistairii. Fungal Planet No. 4.<br />

(www.fungalplanet.org).<br />

Crous PW, Groenewald JZ (2006b). Mycosphaerella maxii. Fungal Planet No. 6.<br />

(www.fungalplanet.org).<br />

Crous PW, Groenewald JZ, Gams W (2003). Eyespot of cereals revisited: ITS<br />

phylogeny reveals new species relationships. European Journal of Plant<br />

Pathology 109: 841–850.<br />

Crous PW, Groenewald JZ, Mansilla JP, Hunter GC, Wingfield MJ (2004b).<br />

Phylogenetic reassessment of Mycosphaerella spp. <strong>and</strong> their anamorphs<br />

occurring on Eucalyptus. Studies in Mycology 50: 195–214.<br />

Crous PW, Kang JC, Braun U (2001). A phylogenetic redefinition of anamorph<br />

genera in Mycosphaerella based on ITS rDNA sequence <strong>and</strong> morphology.<br />

Mycologia 93: 1081–1101.<br />

Crous PW, Liebenberg MM, Braun U, Groenewald JZ (2006a). Re-evaluating the<br />

taxonomic status of Phaeoisariopsis griseola, the causal agent of angular leaf<br />

spot of bean. Studies in Mycology 55: 163–173.<br />

Crous PW, Palm ME (1999). Systematics of selected foliicolous fungi associated<br />

with leaf spots of Proteaceae. Mycological Research 103: 1299–1304.<br />

Crous PW, Schroers HJ, Groenewald JZ, Braun U, Schubert K (2006d).<br />

Metulocladosporiella gen. nov. for the causal organism of <strong>Cladosporium</strong><br />

speckle disease of banana. Mycological Research 110: 264–275.<br />

Crous PW, Summerell BA, Carnegie AJ, Mohammed C, Himaman W, Groenewald<br />

JZ (2007b). Foliicolous Mycosphaerella spp. <strong>and</strong> their anamorphs on Corymbia<br />

<strong>and</strong> Eucalyptus. Fungal Diversity 26: 143–185.<br />

Crous PW, Wingfield MJ, Mansilla JP, Alfenas AC, Groenewald JZ (2006b).<br />

Phylogenetic reassessment of Mycosphaerella spp. <strong>and</strong> their anamorphs<br />

occurring on Eucalyptus. II. Studies in Mycology 55: 99–131.<br />

Crous PW, Slippers B, Wingfield MJ, Rheeder J, Marasas WFO, Phillips AJL, Alves<br />

A, Burgess T, Barber P, Groenewald JZ (2006c). Phylogenetic lineages in the<br />

Botryosphaeriaceae. Studies in Mycology 55: 235–253.<br />

Doherty MW (1900). New Species of Trimmatostroma. Botanical Gazette 30: 400–<br />

403.<br />

Dugan FM, Schubert K, Braun U (2004). Check-list of <strong>Cladosporium</strong> names.<br />

Schlechtendalia 11: 1–103.<br />

Ellis MB (1976). More <strong>dematiaceous</strong> hyphomycetes. Commonwealth Mycological<br />

Institute, Kew.<br />

Eriksson O (1981). <strong>The</strong> families of bitunicate ascomycetes. Opera Botanica 60:<br />

1–220.<br />

www.studiesinmycology.org<br />

31


Crous et al.<br />

Gams W, Verkley GJM, Crous PW (2007). <strong>CBS</strong> course of mycology, 5 th ed.<br />

Centraalbureau voor Schimmelcultures, Utrecht.<br />

Goodwin SB, Dunkley LD, Zismann VL (2001). Phylogenetic analysis of Cercospora<br />

<strong>and</strong> Mycosphaerella based on the internal transcribed spacer region of<br />

ribosomal DNA. Phytopathology 91: 648–658.<br />

Gorbushina AA, Panina LK, Vlasov DY, Krumbein WE (1996). Fungi deteriorating<br />

Chersonessus marbles. Mikologia i Fitopatologia 30: 23–28.<br />

Hambleton S, Tsuneda A, Currah RS (2003). Comparative morphology <strong>and</strong><br />

phylogenetic placement of two microsclerotial black fungi from Sphagnum.<br />

Mycologia 95: 959–975.<br />

Hoog GS de, Gerrits van den Ende AHG (1998). Molecular diagnostics of clinical<br />

strains of filamentous Basidiomycetes. Mycoses 41: 183–189.<br />

Hoog GS de, Guarro J, Gené J, Figueras MJ (2000). Atlas of Clinical Fungi. 2 nd<br />

Edn. Centraalbureau voor Schimmelcultures, Utrecht, <strong>and</strong> Universitat Rovira<br />

I Virgili, Reus.<br />

Hoog GS de, Oorschot CAN van, Hijwegen T (1983). Taxonomy of the Dactylaria<br />

complex. II. Proceedings van de Koninklijke Nederl<strong>and</strong>se Akademie van<br />

Wetenschappen, Series C, 86(2): 197–206.<br />

Hoog GS de, Zalar P, Urzì C, Leo de F, Yurlova NA, Sterflinger K (1999). Relationships<br />

of dothideaceous black yeasts <strong>and</strong> meristematic fungi based on 5.8S <strong>and</strong> ITS2<br />

rDNA sequence comparision. Studies in Mycology 43: 31–37.<br />

Hunter GC, Wingfield BD, Crous PW, Wingfield MJ (2006). A multi-gene phylogeny<br />

for species of Mycosphaerella occurring on Eucalyptus leaves. Studies in<br />

Mycology 55: 147–161.<br />

Kirk PM (1981). New or interesting microfungi. 1. Dematiaceous hyphomycetes from<br />

Devon. Transactions of the British Mycological Society 76: 71–87.<br />

Kogej T, Gorbushina AA, Gunde-Cimerman N (2006). Hypersaline conditions<br />

induce changes in cell-wall melanization <strong>and</strong> colony structure in a halophilic<br />

<strong>and</strong> a xerophilic black yeast species of the <strong>genus</strong> Trimmatostroma. Mycological<br />

Research 110: 713–724.<br />

Krumbein WE, Gorbushina AA, Sterflinger K, Haroska U, Kunert U, Drewello R,<br />

Weißmann R (1996). Biodetorioration of historical window panels of the<br />

former Cistercian Monastery church of Haina (Hessen, Germany). DECHEMA<br />

Monographs 133: 417–424.<br />

Moncalvo J-M, Rehner SA, Vilgalys R (1993). Systematics of Lyophyllum section<br />

Difformia based on evidence from culture studies <strong>and</strong> ribosomal DNA<br />

sequences. Mycologia 85: 788–794.<br />

Müller E, Oehrens E (1982). On the <strong>genus</strong> Teratosphaeria (Ascomycetes). Sydowia<br />

35: 138–142.<br />

Niekerk JM van, Groenewald JZ, Verkley GJM, Fourie PH, Wingfield MJ, Crous PW<br />

(2004). Systematic reappraisal of Coniella <strong>and</strong> Pilidiella, with specific reference<br />

to species occurring on Eucalyptus <strong>and</strong> Vitis in South Africa. Mycological<br />

Research 108: 283–303.<br />

Rayner RW (1970). A mycological colour chart. CMI <strong>and</strong> British Mycological Society.<br />

Kew.<br />

Rehner SA, Samuels GJ (1994). Taxonomy <strong>and</strong> phylogeny of Gliocladium analysed<br />

from nuclear large subunit ribosomal DNA sequences. Mycological Research<br />

98: 625–634.<br />

Roux C, Triebel D (1994). Révision des espèces de Stigmidium et de<br />

Sphaerellothecium (champignons lichénicoles non lichénisés, Ascomycetes)<br />

correspondant à Pharcidia epicymatia sensu Keissler ou à Stigmidium schaereri<br />

auct. Bulletin de la Societe Linneenne de Provence 45: 451–542.<br />

Schoch C, Shoemaker RA, Seifert, KA, Hambleton S, Spatafora JW, Crous PW<br />

(2006). A multigene phylogeny of the Dothideomycetes using four nuclear loci.<br />

Mycologia 98: 1043–1054.<br />

Schubert K, Groenewald JZ, Braun U, Dijksterhuis J, Starink M, Hill CF, Zalar P,<br />

Hoog GS de, Crous PW (2007). Biodiversity in the <strong>Cladosporium</strong> herbarum<br />

complex (Davidiellaceae, Capnodiales), with st<strong>and</strong>ardisation of methods for<br />

<strong>Cladosporium</strong> taxonomy <strong>and</strong> diagnostics. Studies in Mycology 58: 105–156.<br />

Seifert KA, Nickerson NL, Corlett M, Jackson ED, Louis-Seize G, Davies RJ<br />

(2004). Devriesia, a new hyphomycete <strong>genus</strong> to accommodate heat-resistant,<br />

cladosporium-like fungi. Canadian Journal of Botany 82: 914–926.<br />

Selbmann L, Hoog GS de, Mazzaglia A, Friedmann EI, Onofri S (2005). Fungi at<br />

the edge of life: cryptoendolithic black fungi from Antarctic desert. Studies in<br />

Mycology 51: 1–32.<br />

Stewart EL, Liu Z, Crous PW, Szabo L (1999). Phylogenetic relationships among<br />

some cercosporoid anamorphs of Mycosphaerella based on rDNA sequence<br />

analysis. Mycological Research 103: 1491–1499.<br />

Sugiyama J, Amano N (1987). Two metacapnodiaceous sooty moulds from Japan:<br />

their identity <strong>and</strong> behavior in pure culture. In: Pleomorphic fungi: the diversity<br />

<strong>and</strong> its taxonomic implications. (Sugiyama J, ed.). Kodansha, Tokyo/Elsevier,<br />

Amsterdam: 141–156.<br />

Summerell BA, Groenewald JZ, Carnegie AJ, Summerbell RC, Crous PW (2006).<br />

Eucalyptus microfungi known from culture. 2. Alysidiella, Fusculina <strong>and</strong><br />

Phlogicylindrium genera nova, with notes on some other poorly known taxa.<br />

Fungal Diversity 23: 323–350.<br />

Sutton BC (1971). Staninwardia gen. nov. (Melanconiales) on Eucalyptus.<br />

Transactions of the British Mycological Society 57: 539–542.<br />

Sutton BC, Pascoe IG (1989). Reassessment of Peltosoma, Stigmina <strong>and</strong><br />

Batcheloromyces <strong>and</strong> description of Hyphothyrium gen. nov. Mycological<br />

Research 92: 210–222.<br />

Taylor JE, Crous PW, Wingfield MJ (1999). <strong>The</strong> <strong>genus</strong> Batcheloromyces on<br />

Proteaceae. Mycological Research 103: 1478–1484.<br />

Taylor JE, Groenewald JZ, Crous PW (2003). A phylogenetic analysis of<br />

Mycosphaerellaceae leaf spot pathogens of Proteaceae. Mycological Research<br />

107: 653–658.<br />

Verkley GJM, Crous PW, Groenewald JZ, Braun U, Aptroot A (2004). Mycosphaerella<br />

punctiformis revisited: morphology, phylogeny, <strong>and</strong> epitypification of the type<br />

species of the <strong>genus</strong> Mycosphaerella (Dothideales, Ascomycota). Mycological<br />

Research 108: 1271–1282.<br />

Vilgalys R, Hester M (1990). Rapid genetic identification <strong>and</strong> mapping of<br />

enzymatically amplified ribosomal DNA from several Cryptococcus species.<br />

Journal of Bacteriology 172: 4238–4246.<br />

White TJ, Bruns T, Lee S, Taylor J (1990). Amplification <strong>and</strong> direct sequencing of<br />

fungal ribosomal RNA genes for phylogenetics. In: PCR Protocols: a guide to<br />

methods <strong>and</strong> applications (Innis MA, Gelf<strong>and</strong> DH, Sninsky JJ, White TJ, eds).<br />

Academic Press, San Diego, California: 315–322.<br />

Winka K, Eriksson OE, Bång Å (1998). Molecular evidence for recognizing the<br />

Chaetothyriales. Mycologia 90: 822–830.<br />

Wollenzien U, Hoog GS de, Krumbein WE, Urzì C (1995). On the isolation of<br />

microcolonial fungi occurring on <strong>and</strong> in marble <strong>and</strong> other calcareous rocks.<br />

Science of the Total Environment 167: 287–294.<br />

32


available online at www.studiesinmycology.org<br />

doi:10.3114/sim.2007.58.02<br />

Studies in Mycology 58: 33–56. 2007.<br />

Delimiting <strong>Cladosporium</strong> from morphologically <strong>similar</strong> genera<br />

P.W. Crous 1* , U. Braun 2 , K. Schubert 3 <strong>and</strong> J.Z. Groenewald 1<br />

1<br />

<strong>CBS</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; 2 Martin-Luther-Universität, Institut für Biologie, Geobotanik und Botanischer Garten,<br />

Herbarium, Neuwerk 21, D-06099 Halle (Saale), Germany; 3 Botanische Staatssammlung München, Menzinger Straße 67, D-80638 München, Germany<br />

*Correspondence: Pedro Crous, p.crous@cbs.knaw.nl<br />

Abstract: <strong>The</strong> <strong>genus</strong> <strong>Cladosporium</strong> is restricted to <strong>dematiaceous</strong> hyphomycetes with a coronate scar type, <strong>and</strong> Davidiella teleomorphs. In the present study numerous<br />

cladosporium-like taxa are treated, <strong>and</strong> allocated to different genera based on their morphology <strong>and</strong> DNA phylogeny derived from the LSU nrRNA gene. Several species are<br />

introduced in new genera such as Hyalodendriella, Ochrocladosporium, Rachicladosporium, Rhizocladosporium, Toxicocladosporium <strong>and</strong> Verrucocladosporium. A further new<br />

taxon is described in Devriesia (Teratosphaeriaceae). Furthermore, <strong>Cladosporium</strong> castellanii, the etiological agent of tinea nigra in humans, is confirmed as synonym of Stenella<br />

araguata, while the type species of Stenella is shown to be linked to the Teratosphaeriaceae (Capnodiales), <strong>and</strong> not the Mycosphaerellaceae as formerly presumed.<br />

Taxonomic novelties: Devriesia americana Crous & Dugan, sp. nov., Hyalodendriella Crous, gen. nov., Hyalodendriella betulae Crous sp. nov., Ochrocladosporium Crous &<br />

U. Braun, gen. nov., Ochrocladosporium elatum (Harz) Crous & U. Braun, comb. nov., Ochrocladosporium frigidarii Crous & U. Braun, sp. nov., Rachicladosporium Crous, U.<br />

Braun & Hill, gen. nov., Rachicladosporium luculiae Crous, U. Braun & Hill, sp. nov., Rhizocladosporium Crous & U. Braun, gen. nov., Rhizocladosporium argillaceum (Minoura)<br />

Crous & U. Braun, comb. nov., Toxicocladosporium Crous & U. Braun, gen. nov., Toxicocladosporium irritans Crous & U. Braun, sp. nov., Verrucocladosporium K. Schub., Aptroot<br />

& Crous, gen. nov., Verrucocladosporium dirinae K. Schub., Aptroot & Crous, sp. nov.<br />

Key words: <strong>Cladosporium</strong>, Davidiella, food spoilage, hyphomycetes, indoor air, LSU phylogeny, taxonomy.<br />

Introduction<br />

Cladosporioid hyphomycetes are common, widespread fungi.<br />

<strong>The</strong> <strong>genus</strong> <strong>Cladosporium</strong> Link is based on the type species,<br />

<strong>Cladosporium</strong> herbarum (Pers. : Fr.) Link, which in turn has been<br />

linked to Davidiella Crous & U. Braun teleomorphs (Braun et al.<br />

2003, Schubert et al. 2007b – this volume). <strong>Cladosporium</strong> is one<br />

of the largest, most heterogeneous genera of hyphomycetes,<br />

comprising more than 772 names (Dugan et al. 2004), <strong>and</strong> including<br />

endophytic, fungicolous, human pathogenic, phytopathogenic <strong>and</strong><br />

saprobic species. Species of this <strong>genus</strong> affect daily human life in<br />

various ways. <strong>The</strong> common saprobic members of <strong>Cladosporium</strong><br />

occur on all kinds of senescing <strong>and</strong> dead leaves <strong>and</strong> stems of<br />

herbaceous <strong>and</strong> woody plants, as secondary invaders on necrotic<br />

leaf lesions caused by other fungi, are frequently isolated from air,<br />

soil, food stuffs, paint, textiles <strong>and</strong> other organic matters, are also<br />

known to be common endophytes (Riesen & Sieber 1985, Brown<br />

et al. 1998, El-Morsy 2000) as well as phylloplane fungi (Islam &<br />

Hasin 2000, De Jager et al. 2001, Inacio et al. 2002, Stohr & Dighton<br />

2004, Levetin & Dorsey 2006). Furthermore, some <strong>Cladosporium</strong><br />

species are known to be potential agents of medical relevance.<br />

<strong>Cladosporium</strong> herbarum is, for instance, a common contaminant in<br />

clinical laboratories <strong>and</strong> causes allergic lung mycoses (de Hoog et<br />

al. 2000, Schubert et al. 2007b – this volume).<br />

In spite of the enormous relevance of this <strong>genus</strong>, there is<br />

no comprehensive modern revision of <strong>Cladosporium</strong>, but some<br />

attempts to revise <strong>and</strong> monograph parts of it have been initiated<br />

during the last decade (David 1997, Partridge & Morgan-Jones<br />

2002, Wirsel et al. 2002, Braun et al. 2003, Dugan et al. 2004, Park<br />

et al. 2004, Seifert et al. 2004, Schubert & Braun 2004, 2005a,<br />

b, 2007, Heuchert et al. 2005, Schubert 2005a, b, Schubert et al.<br />

2006).<br />

Previous molecular studies employing rDNA ITS sequence<br />

data (Crous et al. 2001) have shown <strong>Cladosporium</strong> spp. to cluster<br />

adjacent to the main monophyletic Mycosphaerella Johanson<br />

cluster, suggesting a position apart from the latter <strong>genus</strong>. Braun<br />

et al. (2003) carried out more comprehensive sequence analyses,<br />

based on ITS (ITS-1, 5.8S <strong>and</strong> ITS-2) <strong>and</strong> 18S rDNA data, providing<br />

further evidence that <strong>Cladosporium</strong> s. str. represents a sister clade<br />

of Mycosphaerella.<br />

Various authors discussed the taxonomy <strong>and</strong> circumscription<br />

of <strong>Cladosporium</strong> (von Arx 1981, 1983, McKemy & Morgan-Jones<br />

1990, Braun 1995), reaching different conclusions. However, a<br />

first decisive revision of <strong>Cladosporium</strong>, leading to a more natural<br />

concept of this <strong>genus</strong>, was published by David (1997), who carried<br />

out comprehensive scanning electron microscopic examinations of<br />

the scar <strong>and</strong> hilum structure in <strong>Cladosporium</strong> <strong>and</strong> Heterosporium<br />

Klotzsch ex Cook. <strong>The</strong> first Scanning Electron Micrograph (SEM)<br />

studies of these structures, published by Roquebert (1981), indicated<br />

that the conidiogenous loci <strong>and</strong> conidial hila in <strong>Cladosporium</strong> are<br />

characterised by having a unique structure. David (1997) confirmed<br />

these observations, based on a wide range of <strong>Cladosporium</strong> <strong>and</strong><br />

Heterosporium species, <strong>and</strong> demonstrated that the structures of the<br />

conidiogenous loci <strong>and</strong> hila in the latter <strong>genus</strong> fully agree with those<br />

of <strong>Cladosporium</strong>, proving that Heterosporium was indeed a synonym<br />

of <strong>Cladosporium</strong> s. str. He introduced the term “coronate” for the<br />

<strong>Cladosporium</strong> scar type, which is characterised by having a central<br />

convex part (dome), surrounded by a raised periclinal rim (David<br />

1997), <strong>and</strong> showed that this type is confined to anamorphs, as far<br />

as experimentally proven, connected with teleomorphs belonging<br />

in “Mycosphaerella” s. lat. <strong>The</strong>se results were confirmed in a later<br />

phylogenetic study by Braun et al. (2003). <strong>Cladosporium</strong> s. str. was<br />

shown to be a sister clade to Mycosphaerella s. str., for which the<br />

new teleomorph <strong>genus</strong> Davidiella was proposed. Although no clear<br />

morphological differences were reported between Davidiella <strong>and</strong><br />

Mycosphaerella, a further study by Aptroot (2006) found ascospores<br />

of Davidiella to have characteristic irregular cellular inclusions<br />

(lumina), which are absent in species of Mycosphaerella, along with<br />

periphysoids <strong>and</strong> pseudoparaphyses (Schubert et al. 2007b – this<br />

volume). Furthermore, a higher order phylogeny study by Schoch<br />

et al. (2006), which employed DNA sequence data of four loci (SSU<br />

nrDNA, LSU nrDNA, EF-1α, RPB2), revealed species of Davidiella<br />

to cluster in a separate family (Davidiellaceae) from species of<br />

Mycosphaerella (Mycosphaerellaceae), with both families residing<br />

in the Capnodiales (Dothideomycetes), <strong>and</strong> not Dothideales as<br />

always presumed.<br />

33


Crous et al.<br />

Table 1. Isolates for which new sequences were generated.<br />

Anamorph Teleomorph Accession number 1 Host Country Collector GenBank numbers 2<br />

(ITS, LSU)<br />

Cladoriella eucalypti <strong>CBS</strong> 115899*; CPC 10954 Eucalyptus sp. South Africa P.W. Crous EU040224, EU040224<br />

Coniothyrium palmarum <strong>CBS</strong> 758.73; CMW 5283 Phoenix dactylifera Israel Y. Pinkas DQ240000, EU040225<br />

Devriesia acadiensis <strong>CBS</strong> 115874; DAOM 232211 Soil Canada N. Nickerson AY692095, EU040226<br />

Devriesia americana <strong>CBS</strong> 117726; ATCC 96545; CPC 5121 Air U.S.A. F.M. Dugan AY251068, EU040227<br />

Devriesia shelburniensis <strong>CBS</strong> 115876; DAOM 232217 Soil Canada N. Nickerson AY692093, EU040228<br />

Devriesia thermodurans <strong>CBS</strong> 115878*; DAOM 225330 Soil Canada N. Nickerson AY692087, EU040229<br />

Hormoconis resinae <strong>CBS</strong> 365.86 – – – EU040230, EU040230<br />

<strong>CBS</strong> 184.54; ATCC 11841; CPC 3692; IMI 089837; IFO 31706 Creosote-treated wooden pole U.S.A. – AY251067, EU040231<br />

Hyalodendriella betulae <strong>CBS</strong> 261.82* Alnus glutinosa Netherl<strong>and</strong>s W. Gams EU040232, EU040232<br />

Ochrocladosporium elatum <strong>CBS</strong> 146.33*; ATCC 11280; ATHUM 2862; IFO 6372; IMI 049629; Wood pulp Sweden E. Melin EU040233, EU040233<br />

MUCL 10094<br />

Ochrocladosporium frigidarii <strong>CBS</strong> 103.81* Cooled room Germany B. Ahlert EU040234, EU040234<br />

Parapleurotheciopsis inaequiseptata MUCL 41089; INIFAT C98/30-1 Rotten leaf Brazil R.F. Castañeda EU040235, EU040235<br />

Passalora daleae <strong>CBS</strong> 113031* Dalea spinosa Mexico L.B. Sparrius EU040236, EU040236<br />

Rachicladosporium luculiae CPC 11407* Luculia sp. New Zeal<strong>and</strong> F. Hill EU040237, EU040237<br />

Ramularia aplospora Mycosphaerella alchemillicola <strong>CBS</strong> 545.82* Powdery mildew on Alchemilla vulgaris Germany T. Hijwegen EU040238, EU040238<br />

Retroconis fusiformis <strong>CBS</strong> 330.81; IMI 170799 Gossypium sp. Pakistan – EU040239, EU040239<br />

Rhizocladosporium argillaceum <strong>CBS</strong> 241.67*; ATCC 38103; IFO 7055; OUT 4262 Decayed myxomycete Japan K. Tubaki EU040240, EU040240<br />

Subramaniomyces fusisaprophyticus <strong>CBS</strong> 418.95; INIFAT C94/134 Leaf litter Cuba R.F. Castañeda EU040241, EU040241<br />

<strong>The</strong>dgonia ligustrina W1877 Ligustrum sp. – H. Evans EU040242, EU040242<br />

Toxicocladosporium irritans <strong>CBS</strong> 185.58* Mouldy paint Suriname M.B. Schol-Schwarz EU040243, EU040243<br />

Verrucocladosporium dirinae <strong>CBS</strong> 112794* Dirina massiliensis U.K. A. Aptroot EU040244, EU040244<br />

1 ATCC: American Type Culture Collection, Virginia, U.S.A.; ATHUM: Culture Collection of Fungi, University of Athens, Department of Biology, Section of Ecology <strong>and</strong> Systematics, Athens, Greece; <strong>CBS</strong>: Centraalbureau voor Schimmelcultures, Utrecht,<br />

<strong>The</strong> Netherl<strong>and</strong>s; CMW: Culture collection of Mike Wingfield, housed at FABI, Pretoria, South Africa; CPC: Culture collection of Pedro Crous, housed at <strong>CBS</strong>; DAOM: Plant Research Institute, Department of Agriculture (Mycology), Ottawa, Canada; IFO:<br />

Institute For Fermentation, Osaka, Japan; IMI: International Mycological Institute, CABI-Bioscience, Egham, Bakeham Lane, U.K.; INIFAT: Alex<strong>and</strong>er Humboldt Institute for Basic Research in Tropical Agriculture, Ciudad de La Habana, Cuba; MUCL:<br />

Mycotheque de l’ Université Catholique de Louvain, Louvain-la-Neuve, Belgium; OUT: Department of Fermentation Technology, Faculty of Engineering, Osaka University, Yamadaue, Suita-shi, Osaka, Japan.<br />

2 ITS: internal transcribed spacer regions <strong>and</strong> 5.8S rRNA gene; LSU: partial 28S rRNA gene.<br />

*Ex-type cultures.<br />

34


<strong>Cladosporium</strong> <strong>and</strong> morphologically <strong>similar</strong> genera<br />

<strong>The</strong> current circumscription of <strong>Cladosporium</strong> emend. can be<br />

summarised as follows: Dematiaceous hyphomycetes; Davidiella<br />

anamorphs; mycelium internal <strong>and</strong> external; hyphae branched,<br />

septate, pigmented; stromata lacking or occasionally present;<br />

conidiophores mononematous, solitary to fasciculate, cylindrical,<br />

geniculate-sinuous to nodulose, simple to branched, subhyaline<br />

to usually distinctly pigmented, continuous to septate, smooth<br />

to verruculose; conidiogenous cells integrated, terminal <strong>and</strong><br />

intercalary, usually sympodial, with a single to several scars;<br />

conidiogenesis holoblastic; conidiogenous loci coronate, i.e.,<br />

more or less protuberant, composed of a central convex dome,<br />

surrounded by a raised periclinal rim, barely to distinctly darkened;<br />

conidia solitary or in short to long, simple to branched acropetal<br />

chains, amero- to phragmosporous, subhyaline to usually distinctly<br />

pigmented, smooth, verruculose, verrucose, echinulate, cristate,<br />

hila coronate, more or less protuberant.<br />

<strong>The</strong> new concept of <strong>Cladosporium</strong> s. str., supported by<br />

molecular data <strong>and</strong> typical coronate conidiogenous loci <strong>and</strong> conidial<br />

hila, rendered it possible to initiate a comprehensive revision<br />

of <strong>Cladosporium</strong> s. lat. <strong>The</strong> preparation of a general, annotated<br />

check-list of <strong>Cladosporium</strong> s. lat. was the first step in this direction<br />

(Dugan et al. 2004). <strong>The</strong> aim of the present study, therefore, was to<br />

delineate <strong>Cladosporium</strong> s. str. from other taxa that have in recent<br />

years been described in <strong>Cladosporium</strong> s. lat. To attain this goal<br />

isolates were studied under st<strong>and</strong>ardised conditions on a set of<br />

predescribed media (Schubert et al. 2007b – this volume), <strong>and</strong><br />

subjected to DNA sequence analysis of the LSU nrRNA gene.<br />

were only sequenced for isolates of which these data were not<br />

available. <strong>The</strong> ITS data were not included in the analyses but<br />

deposited in GenBank where applicable. Gaps longer than 10<br />

bases were coded as single events for the phylogenetic analyses;<br />

the remaining gaps were treated as missing data. Sequence data<br />

were deposited in GenBank (Table 1) <strong>and</strong> alignments in TreeBASE<br />

(www.treebase.org).<br />

Morphology<br />

Wherever possible, 30 measurements (× 1 000 magnification)<br />

were made of structures mounted in lactic acid or Shear’s solution<br />

(Gams et al. 2007), with the extremes of spore measurements<br />

given in parentheses. Microscopic observations were made from<br />

colonies cultivated for 7 d under continuous near-ultraviolet light<br />

at 25 °C on SNA as explained in Schubert et al. (2007b – this<br />

volume). Three classes of conidia are distinguished. Ramoconidia<br />

are defined as short apical branches (often conidiogenous cells)<br />

of a conidiophore which secede <strong>and</strong> function as conidia. <strong>The</strong>y are<br />

characterised by having a truncate, undifferentiated base, i.e.,<br />

they differ from true conidia by lacking characteristic basal hila<br />

caused by conidiogenesis. Ramoconidia give rise to branched or<br />

unbranched conidia. Secondary ramoconidia are branched conidia<br />

with a narrowed base, bearing a true hilum, that can occur in chains,<br />

giving rise to conidia, which differ from secondary ramoconidia<br />

with regards to shape, size <strong>and</strong> septation. In previous literature on<br />

<strong>Cladosporium</strong> <strong>and</strong> allied genera, the true ramoconidia have often<br />

been classified as “ramoconidia s. str.” whereas the secondary<br />

ramoconidia have been named “ramoconidia s. lat.”<br />

Materials <strong>and</strong> methods<br />

Isolates<br />

Isolates used were obtained from the Centraalbureau voor<br />

Schimmelcultures (<strong>CBS</strong>), or freshly isolated from various<br />

substrates (Table 1). Strains were cultured on 2 % malt extract<br />

plates (MEA; Gams et al. 2007), by obtaining single conidial<br />

colonies as explained in Crous (1998). Colonies were subcultured<br />

onto fresh MEA, oatmeal agar (OA), potato-dextrose agar (PDA)<br />

<strong>and</strong> synthetic nutrient-poor agar (SNA) (Gams et al. 2007), <strong>and</strong><br />

incubated under near-ultraviolet light to study their morphology.<br />

Cultural characteristics were assessed after 2–4 wk on OA <strong>and</strong><br />

PDA at 25 °C in the dark, using the colour charts of Rayner (1970).<br />

Nomenclatural novelties <strong>and</strong> descriptions were deposited in<br />

MycoBank (www.MycoBank.org).<br />

DNA isolation, sequencing <strong>and</strong> phylogeny<br />

Fungal colonies were established on agar plates, <strong>and</strong> genomic<br />

DNA was isolated following the CTAB-based protocol described in<br />

Gams et al. (2007). <strong>The</strong> primers V9G (de Hoog & Gerrits van den<br />

Ende 1998) <strong>and</strong> LR5 (Vilgalys & Hester 1990) were used to amplify<br />

part of the nuclear rDNA operon spanning the 3’ end of the 18S<br />

rRNA gene (SSU), the first internal transcribed spacer (ITS1), the<br />

5.8S rRNA gene, the second ITS region <strong>and</strong> the 5’ end of the 28S<br />

rRNA gene (LSU). Four internal primers, namely ITS4 (White et al.<br />

1990), LR0R (Rehner & Samuels 1994), LR3R (www.biology.duke.<br />

edu/fungi/mycolab/primers.htm), <strong>and</strong> LR16 (Moncalvo et al. 1993),<br />

were used for sequencing to ensure that good quality overlapping<br />

sequences are obtained. <strong>The</strong> PCR conditions, sequence alignment<br />

<strong>and</strong> subsequent phylogenetic analysis followed the methods of<br />

Crous et al. (2006d). <strong>The</strong> ITS1, ITS2 <strong>and</strong> 5.8S rRNA gene (ITS)<br />

www.studiesinmycology.org<br />

Results<br />

DNA extraction, amplification <strong>and</strong> phylogeny<br />

Amplicons of approximately 1 700 bases were obtained for the<br />

isolates listed in Table 1. <strong>The</strong> newly generated sequences were<br />

used to obtain additional sequences from GenBank, which were<br />

added to the alignment. <strong>The</strong> manually adjusted LSU alignment<br />

contained 73 sequences (including the two outgroup sequences)<br />

<strong>and</strong> 996 characters including alignment gaps. Of the 849 characters<br />

used in the phylogenetic analysis, 336 were parsimony-informative,<br />

77 were variable <strong>and</strong> parsimony-uninformative, <strong>and</strong> 436 were<br />

constant. Neighbour-joining analyses using three substitution<br />

models on the sequence data yielded trees with identical topologies<br />

to one another. <strong>The</strong> neighbour-joining trees support the same<br />

clades as obtained from the parsimony analysis, but with a different<br />

arrangement at the deeper nodes, which were poorly supported<br />

in the bootstrap analyses or not at all (for example, the Helotiales<br />

<strong>and</strong> Pleosporales are swapped around). Performing a parsimony<br />

analysis with gaps treated as new characters increases the number<br />

of equally parsimonious trees to 94; the same topology is observed<br />

but with less resolution for the taxa in the Helotiales (data not<br />

shown). Forty-four equally most parsimonious trees (TL = 1 572<br />

steps; CI = 0.436; RI = 0.789; RC = 0.344), one of which is shown<br />

in Fig. 1, were obtained from the parsimony analysis of the LSU<br />

sequence data. <strong>The</strong> cladosporium-like taxa were found to belong to<br />

the Helotiales, Pleosporales, Sordariales <strong>and</strong> as sister taxa to the<br />

Davidiellaceae in the Capnodiales.<br />

<strong>The</strong> LSU alignment used for parsimony <strong>and</strong> distance analysis<br />

was supplemented with sequences for Parapleurotheciopsis<br />

inaequiseptata (Matsush.) P.M. Kirk <strong>and</strong> Subramaniomyces<br />

fusisaprophyticus (Matsush.) P.M. Kirk, as well as related sequences<br />

35


Crous et al.<br />

100<br />

61<br />

100<br />

52<br />

100<br />

10 changes<br />

Athelia epiphylla AY586633<br />

Paullicorticium ansatum AY586693<br />

Rhizocladosporium argillaceum <strong>CBS</strong> 241.67<br />

Myxotrichum deflexum AY541491<br />

Hormoconis resinae <strong>CBS</strong> 365.86<br />

Hormoconis resinae <strong>CBS</strong> 184.54<br />

56<br />

99<br />

99<br />

67<br />

100 84<br />

90<br />

93<br />

81<br />

100<br />

65<br />

75<br />

100<br />

Bisporella citrina AY789385<br />

Sarcoleotia turficola AY789277<br />

Torrendiella eucalypti DQ195799<br />

Torrendiella eucalypti DQ195800<br />

Hyalodendriella betulae <strong>CBS</strong> 261.82<br />

<strong>The</strong>dgonia ligustrina W1877<br />

Blumeria graminis f. sp. bromi AB022362 Erysiphaceae<br />

Neofabraea alba AY064705 Dermateaceae<br />

Neofabraea malicorticis AY544662<br />

Phoma herbarum DQ678066<br />

Ascochyta pisi DQ678070<br />

Didymella bryoniae AB266850<br />

Didymella cucurbitacearum AY293792<br />

Leptospora rubella DQ195792<br />

Phaeosphaeria avenaria AY544684<br />

Coniothyrium palmarum <strong>CBS</strong> 758.73<br />

Ochrocladosporium frigidarii <strong>CBS</strong> 103.81<br />

Ochrocladosporium elatum <strong>CBS</strong> 146.33<br />

Cladoriella eucalypti DQ195790<br />

100 Cladoriella eucalypti <strong>CBS</strong> 115899<br />

100<br />

83<br />

87<br />

85<br />

69<br />

Helotiaceae<br />

Chaetomium homopilatum AF286404<br />

Retroconis fusiformis <strong>CBS</strong> 330.81<br />

Chaetomium globosum AF286403<br />

Aporothielavia leptoderma AF096186<br />

Rachicladosporium luculiae CPC 11407<br />

Dichocladosporium chlorocephalum EU009456<br />

Dichocladosporium chlorocephalum EU009458<br />

Dichocladosporium chlorocephalum EU009457<br />

Dichocladosporium chlorocephalum EU009455<br />

Verrucocladosporium dirinae <strong>CBS</strong> 112794<br />

Toxicocladosporium irritans <strong>CBS</strong> 185.58<br />

97<br />

93<br />

89<br />

100<br />

<strong>Cladosporium</strong> cladosporioides DQ008145<br />

<strong>Cladosporium</strong> uredinicola DQ008147<br />

<strong>Cladosporium</strong> bruhnei DQ008149<br />

<strong>Cladosporium</strong> iridis DQ008148<br />

<strong>Cladosporium</strong> cladosporioides DQ008146<br />

Mycosphaerella marksii AF309578<br />

Penidiella nect<strong>and</strong>rae EU019275<br />

“Stenella” cerophilum AF050286<br />

84<br />

55<br />

100<br />

75<br />

94<br />

100<br />

99<br />

100<br />

94<br />

58<br />

52<br />

58<br />

Incertae sedis<br />

Myxotrichaceae<br />

Amorphotecaceae<br />

Incertae sedis<br />

Incertae sedis<br />

“Mycosphaerella” lateralis AF309583<br />

Pseudocercospora paraguayensis AF309574<br />

Passalora eucalypti AF309575<br />

Mycosphaerella irregulariramosa AF309576<br />

Mycosphaerella punctiformis AY490776<br />

Ramularia aplospora <strong>CBS</strong> 545.82<br />

Passalora daleae <strong>CBS</strong> 113031<br />

Mycosphaerella africana AF309581<br />

Passalora fulva DQ008163<br />

Staninwardia suttonii DQ923535<br />

Devriesia americana <strong>CBS</strong> 117726<br />

Penidiella strumelloidea EU019277<br />

Devriesia thermodurans <strong>CBS</strong> 115878<br />

64<br />

66<br />

Devriesia shelburniensis <strong>CBS</strong> 115876<br />

Devriesia acadiensis <strong>CBS</strong> 115874<br />

Devriesia staurophora DQ008150<br />

Devriesia staurophora DQ008151<br />

Catenulostroma germanicum EU019253<br />

Catenulostroma chromoblastomycosum EU019251<br />

Penidiella rigidophora EU019276<br />

Teratosphaeria alistairii DQ885901<br />

Penidiella columbiana EU019274<br />

Penidiella venezuelensis EU019278<br />

Catenulostroma castellanii EU019250<br />

Teratosphaeria suttonii AF309587<br />

Teratosphaeria molleriana AF309584<br />

Teratosphaeria cryptica AF309585<br />

Teratosphaeria juvenis AF309586<br />

Phaeosphaeriaceae<br />

Leptosphaeriaceae<br />

Incertae sedis<br />

Incertae sedis<br />

Chaetomiaceae, Sordariales<br />

Incertae sedis<br />

Davidiellaceae<br />

Pleosporales<br />

Fig. 1. One of 44 equally most parsimonious trees obtained from a heuristic search with 100 r<strong>and</strong>om taxon additions of the LSU sequence alignment using PAUP v. 4.0b10. <strong>The</strong><br />

scale bar shows 10 changes, <strong>and</strong> bootstrap support values from 1 000 replicates are shown at the nodes. Thickened lines indicate the strict consensus branches <strong>and</strong> ex-type<br />

sequences are printed in bold face. <strong>The</strong> tree was rooted to two sequences obtained from GenBank (Athelia epiphylla AY586633 <strong>and</strong> Paullicorticium ansatum AY586693).<br />

Mycosphaerellaceae<br />

Teratosphaeriaceae<br />

Helotiales<br />

Capnodiales<br />

36


<strong>Cladosporium</strong> <strong>and</strong> morphologically <strong>similar</strong> genera<br />

Athelia epiphylla AY586633<br />

Paullicorticium ansatum AY586693<br />

Blumeria graminis f. sp. bromi AB022362 Erysiphaceae<br />

Bisporella citrina AY789385 Helotiaceae<br />

0.80<br />

Hyalodendriella betulae <strong>CBS</strong> 261.82<br />

<strong>The</strong>dgonia ligustrina W1877<br />

Incertae sedis<br />

1.00<br />

0.69 Neofabraea alba AY064705<br />

Dermateaceae<br />

1.00 Neofabraea malicorticis AY544662<br />

Helotiales<br />

0.60 Sarcoleotia turficola AY789277 Helotiaceae<br />

1.00 Torrendiella eucalypti DQ195799<br />

0.90<br />

Incertae sedis<br />

Torrendiella eucalypti DQ195800<br />

1.00 Hormoconis resinae <strong>CBS</strong> 365.86<br />

Amorphotecaceae<br />

Hormoconis resinae <strong>CBS</strong> 184.54<br />

0.94<br />

Rhizocladosporium argillaceum <strong>CBS</strong> 241.67 Incertae sedis<br />

Myxotrichum deflexum AY541491 Myxotrichaceae<br />

1.00 Chaetomium globosum AF286403<br />

0.88<br />

Aporothielavia leptoderma AF096186<br />

0.95<br />

Chaetomiaceae, Sordariales<br />

1.00 0.79<br />

Chaetomium homopilatum AF286404<br />

Retroconis fusiformis <strong>CBS</strong> 330.81<br />

1.00 Fasciatispora petrakii AY083828 Xylariaceae<br />

0.81<br />

Phlogicylindrium eucalypti DQ923534 Incertae sedis<br />

Plectosphaera eucalypti DQ923538 Phyllachoraceae<br />

0.72<br />

Xylariales<br />

0.74 Subramaniomyces fusisaprophyticus <strong>CBS</strong> 418.95 Incertae sedis<br />

0.68 Parapleurotheciopsis inaequiseptata MUCL 41089 Incertae sedis<br />

Pseudomassaria carolinensis DQ810233 Hyponectriaceae<br />

Phoma herbarum DQ678066<br />

0.63 1.00 Ascochyta pisi DQ678070<br />

1.00 0.97<br />

Incertae sedis<br />

Didymella bryoniae AB266850<br />

1.00 Didymella cucurbitacearum AY293792<br />

Leptospora rubella DQ195792<br />

Pleosporales<br />

1.00<br />

Phaeosphaeriaceae<br />

Phaeosphaeria avenaria AY544684<br />

0.78 Coniothyrium palmarum <strong>CBS</strong> 758.73 Leptosphaeriaceae<br />

0.93 Ochrocladosporium frigidarii <strong>CBS</strong> 103.81<br />

1.00 Ochrocladosporium elatum <strong>CBS</strong> 146.33<br />

Incertae sedis<br />

1.00 Cladoriella eucalypti <strong>CBS</strong> 115899<br />

Incertae sedis<br />

1.00 Cladoriella eucalypti DQ195790<br />

Rachicladosporium luculiae CPC 11407<br />

0.85 Dichocladosporium chlorocephalum EU009456<br />

Dichocladosporium chlorocephalum EU009458<br />

0.68 Dichocladosporium chlorocephalum EU009457 Incertae sedis<br />

Dichocladosporium chlorocephalum EU009455<br />

0.89 Verrucocladosporium dirinae <strong>CBS</strong> 112794<br />

Toxicocladosporium irritans <strong>CBS</strong> 185.58<br />

0.90 <strong>Cladosporium</strong> cladosporioides DQ008145<br />

1.00 <strong>Cladosporium</strong> uredinicola DQ008147<br />

0.97 1.00 <strong>Cladosporium</strong> bruhnei DQ008149<br />

Davidiellaceae<br />

0.61 <strong>Cladosporium</strong> iridis DQ008148<br />

1.00 <strong>Cladosporium</strong> cladosporioides DQ008146<br />

Mycosphaerella marksii AF309578<br />

1.00 Penidiella nect<strong>and</strong>rae EU019275<br />

“Stenella” cerophilum AF050286<br />

0.71 “Mycosphaerella” lateralis AF309583<br />

1.00 Pseudocercospora paraguayensis AF309574<br />

0.99 Passalora eucalypti AF309575<br />

0.51 Mycosphaerella irregulariramosa AF309576<br />

1.00 Mycosphaerella punctiformis AY490776<br />

0.86 Ramularia aplospora <strong>CBS</strong> 545.82<br />

Passalora daleae <strong>CBS</strong> 113031<br />

1.00 Mycosphaerella africana AF309581<br />

1.00 Passalora fulva DQ008163<br />

0.94 1.00 Staninwardia suttonii DQ923535<br />

Devriesia americana <strong>CBS</strong> 117726<br />

Penidiella strumelloidea EU019277<br />

1.00 Devriesia thermodurans <strong>CBS</strong> 115878<br />

1.00<br />

Devriesia shelburniensis <strong>CBS</strong> 115876<br />

0.99 Devriesia acadiensis <strong>CBS</strong> 115874<br />

0.86 Devriesia staurophora DQ008150<br />

Devriesia staurophora DQ008151<br />

0.67<br />

Catenulostroma germanicum EU019253<br />

0.65 Penidiella venezuelensis EU019278<br />

Catenulostroma castellanii EU019250<br />

0.1 expected changes per site 0.73 Teratosphaeria suttonii AF309587<br />

1.00 Teratosphaeria molleriana AF309584<br />

1.00 Teratosphaeria cryptica AF309585<br />

Teratosphaeria juvenis AF309586<br />

1.00 Catenulostroma chromoblastomycosum EU019251<br />

0.64<br />

1.00<br />

Penidiella rigidophora EU019276<br />

Teratosphaeria alistairii DQ885901<br />

Penidiella columbiana EU019274<br />

Mycosphaerellaceae<br />

Teratosphaeriaceae<br />

Capnodiales<br />

Fig. 2. Consensus phylogram (50 % majority rule) of 800 trees resulting from a Bayesian analysis of the LSU sequence alignment using MrBayes v. 3.1.2. Bayesian posterior<br />

probabilities are indicated at the nodes. Ex-type sequences are printed in bold face. <strong>The</strong> tree was rooted to two sequences obtained from GenBank (Athelia epiphylla AY586633<br />

<strong>and</strong> Paullicorticium ansatum AY586693).<br />

www.studiesinmycology.org<br />

37


Crous et al.<br />

Fig. 3. Rachicladosporium luculiae (type material). A–F. Conidiophores with conidial chains, <strong>and</strong> conidiogenous loci aggregated in the upper region. G. Conidia. Scale bar =<br />

10 µm.<br />

from GenBank. This alignment was subjected to a Bayesian<br />

analysis using a general time-reversible (GTR) substitution model<br />

with inverse gamma rates <strong>and</strong> dirichlet base frequencies <strong>and</strong> the<br />

temp value set to 0.5. <strong>The</strong> Markov Chain Monte Carlo (MCMC)<br />

analysis of 4 chains started from a r<strong>and</strong>om tree topology <strong>and</strong> lasted<br />

1 000 000 generations. Trees were saved each 1 000 generations,<br />

resulting in 1 000 trees. Burn-in was set at 200 000 generations<br />

after which the likelihood values were stationary, leaving 800 trees<br />

from which the consensus tree (Fig. 2) <strong>and</strong> posterior probabilities<br />

(PP’s) were calculated. <strong>The</strong> average st<strong>and</strong>ard deviation of split<br />

frequencies was 0.018459 at the end of the run. <strong>The</strong> same overall<br />

topology as that observed using parsimony was obtained, with<br />

the main exception that the Helotiales <strong>and</strong> Pleosporales swapped<br />

around, as observed with the distance analysis.<br />

Taxonomy<br />

<strong>The</strong> present study has delineated several cladosporium-like genera<br />

which are phylogenetically unrelated to, <strong>and</strong> morphologically distinct<br />

from <strong>Cladosporium</strong> s. str. (Davidiellaceae, Capnodiales). <strong>The</strong>se are<br />

treated below:<br />

Capnodiales, incertae sedis<br />

Rachicladosporium Crous, U. Braun & C.F. Hill, gen. nov.<br />

MycoBank MB504430.<br />

Etymology: Named after the apical rachis on conidiophores, <strong>and</strong> its<br />

cladosporium-like appearance.<br />

Differt a Cladosporio conidiophoris cum rachibus terminalibus, locis conidiogenis<br />

inconspicuis vel subconspicuis, margine leviter incrassatis, non fuscatis et non<br />

refractivis, hilis inconspicuis.<br />

Mycelium consisting of branched, septate, smooth, hyaline to<br />

pale brown, thin-walled hyphae. Conidiophores erect, solitary,<br />

macronematous, arising from superficial hyphae, subcylindrical,<br />

straight to somewhat geniculate-sinuous, medium brown,<br />

finely verruculose; basal foot cell without swelling or rhizoids.<br />

Conidiogenous cells integrated, terminal, subcylindrical or tips<br />

slightly swollen, forming an apical rachis, multilocal, loci terminal<br />

<strong>and</strong> lateral, without evident sympodial proliferation (non-geniculate);<br />

conidiogenous loci inconspicuous or subconspicuous by being very<br />

slightly thickened along the rim, but neither darkened nor refractive,<br />

giving rise to simple or branched chains or solitary conidia.<br />

Ramoconidia medium brown, finely verruculose, 0–1-septate,<br />

subcylindrical to narrowly ellipsoid; conidia ellipsoid, pale brown,<br />

0(–1)-septate, smooth to finely verruculose; hila inconspicuous;<br />

secession schizolytic.<br />

Type species: Rachicladosporium luculiae Crous, U. Braun & C.F.<br />

Hill, sp. nov.<br />

38


<strong>Cladosporium</strong> <strong>and</strong> morphologically <strong>similar</strong> genera<br />

Rachicladosporium luculiae Crous, U. Braun & C.F. Hill, sp. nov.<br />

MycoBank MB504431. Fig. 3.<br />

Etymology: Named after its host <strong>genus</strong>, Luculia.<br />

Mycelium ex hyphis ramosis, septatis, levibus, hyalinis vel pallide brunneis, 2–3<br />

µm latis compositum. Conidiophora erecta, solitaria, macronemata, ex hyphis<br />

superficialibis oriunda, subcylindrica, recta to geniculata-sinuosa, ad 60 µm longa et<br />

6 µm lata, 3–6-septata, modice brunnea, subtiliter verruculosa, non crassitunicata, ad<br />

basim non inflatae et non rhizoideae. Cellulae conidiogenae integratae, terminales,<br />

8–15 × 4–5 µm, subcylindricae, apicem versus attenuatae, apice obtuso, rachidi<br />

terminali, locis conidialibus numerosis, 1–2 µm latis, margine leviter incrassatis,<br />

non fuscatis et non refractivis. Conidia catenata vel solitaria. Ramoconidia modice<br />

brunnea, subtile verruculosa, 0–1-septata, subcylindrica vel anguste ellipsoidea,<br />

10–17 × 4–5 µm; conidia secundaria ellipsoidea, pallide brunnea, 0(–1)-septata,<br />

levia vel subtile verruculosa, interdum guttulata, (7–)9–12(–15) × 3(–4) µm; hila<br />

inconspicua.<br />

Mycelium consisting of branched, septate, smooth, thin-walled,<br />

hyaline to pale brown, 2–3 µm wide hyphae. Conidiophores<br />

erect, solitary, macronematous, arising from superficial hyphae,<br />

subcylindrical, straight to somewhat geniculate-sinuous, up to 60 µm<br />

long, <strong>and</strong> 6 µm wide, 3–6-septate, medium brown, finely verruculose,<br />

thin-walled (≤ 1 µm), rarely with a single percurrent proliferation;<br />

basal foot cell without swelling or rhizoids. Conidiogenous cells<br />

integrated, terminal, 8–15 × 4–5 µm, subcylindrical, tapering to an<br />

obtuse apex, occasionally slightly swollen at the tip, without distinct<br />

sympodial proliferation (non-geniculate), forming a rachis, with<br />

several conidiogenous loci, terminal <strong>and</strong> lateral, 1–2 µm wide, nonprotuberant,<br />

quite inconspicuous to subconspicuous, very slightly<br />

thickened along the rim, but not darkened <strong>and</strong> refractive; giving rise<br />

to simple or branched chains or solitary conidia, thin-walled (≤ 0.75<br />

µm). Ramoconidia medium brown, finely verruculose, 0–1-septate,<br />

subcylindrical to narrowly ellipsoid, 10–17 × 4–5 µm; conidia<br />

ellipsoid, pale brown, 0(–1)-septate, smooth to finely verruculose,<br />

at times guttulate, (7–)9–12(–15) × 3(–4) µm; hila inconspicuous,<br />

neither thickened nor darkened-refractive.<br />

Cultural characteristics: Colonies on PDA erumpent, spreading,<br />

with moderate aerial mycelium <strong>and</strong> smooth, even margins; irongrey<br />

in the centre, olivaceous-grey in the outer region (surface);<br />

iron-grey underneath. Colonies reaching 4 cm diam after 1 mo at<br />

25 °C in the dark.<br />

Specimen examined: New Zeal<strong>and</strong>, Auckl<strong>and</strong>, isolated from leaf spots on Luculia<br />

sp. (Rubiaceae), 25 Jul. 2004, F. Hill 1059, holotype <strong>CBS</strong> H-19891, culture ex-type<br />

<strong>CBS</strong> 121620 = CPC 11407.<br />

Notes: Rachicladosporium is morphologically quite distinct from<br />

<strong>Cladosporium</strong> s. str. <strong>and</strong> allied cladosporioid genera by having an<br />

apical conidiophore rachis with inconspicuous to subconspicuous<br />

scars <strong>and</strong> unthickened, not darkened-refractive conidial hila. Due<br />

to the structure of the conidiogenous cells, R. luculiae superficially<br />

resembles species of the tretic <strong>genus</strong> Diplococcium Grove (Ellis<br />

1971, 1976; Goh & Hyde 1998). However, there is no evidence<br />

for a tretic conidiogenesis in R. luculiae. <strong>The</strong> conidia are formed<br />

holoblastically <strong>and</strong> separated by a thin septum. Furthermore,<br />

in Diplococcium the conidiogenous cells are terminal as well as<br />

intercalary, the conidiophores are often branched, <strong>and</strong> branched<br />

conidial chains are lacking or at least less common. Molecular<br />

sequence data about Diplococcium species are not yet available,<br />

though taxa that have been analysed show affinities to the<br />

Pleosporaceae <strong>and</strong> Helotiales (Wang et al., unpubl. data), whereas<br />

Rachicladosporium is allied with the Capnodiales. <strong>The</strong> ecology<br />

of R. luculiae is still unclear, although it has been isolated from<br />

lesions on Luculia sp. Fruiting of this species in vivo has not yet<br />

been observed, <strong>and</strong> its pathogenicity remains unproven.<br />

www.studiesinmycology.org<br />

Toxicocladosporium Crous & U. Braun, gen. nov. MycoBank<br />

MB504426.<br />

Etymology: Named after ample volatile metabolites produced in<br />

culture, <strong>and</strong> cladosporium-like morphology.<br />

Differt a Cladosporio locis conidiogenis denticulatis, incrassatis et fuscatis-refractivis,<br />

sed non coronatis, conidiophoris et conidiis cum septis incrassatis et atrofuscis, et<br />

culturis cum metabolitis volaticis toxicis.<br />

Mycelium consisting of branched, septate, dark brown, finely<br />

verruculose hyphae. Conidiophores solitary, dimorphic, solitary,<br />

macronematous or micronematous, reduced to conidiogenous<br />

cells. Macronematous conidiophores subcylindrical, straight<br />

to geniculate-sinuous, or irregularly curved, unbranched or<br />

branched above, septate, dark brown, finely verruculose, walls<br />

thick, septa dark brown; micronematous conidiophores reduced<br />

to conidiogenous cells, erect, doliiform to subcylindrical, with slight<br />

taper towards the apex. Conidiogenous cells integrated, terminal or<br />

lateral, subcylindrical with slight taper towards apex; proliferating<br />

sympodially with apical loci protruding <strong>and</strong> denticle-like, thickened,<br />

darkened <strong>and</strong> refractive, but not coronate. Conidia catenulate in<br />

branched or unbranched chains, medium to dark brown, thickwalled,<br />

with dark, thick septa, smooth to finely verruculose;<br />

ramoconidia septate, prominently constricted at septa, broadly<br />

ellipsoid to subcylindrical; conidia ellipsoid to ovoid, pale to medium<br />

brown, 0(–1)-septate; hila not coronate, but protruding, thickened,<br />

darkened <strong>and</strong> refractive in ramoconidia, but less obvious in young<br />

conidia.<br />

Type species: Toxicocladosporium irritans Crous & U. Braun, sp.<br />

nov.<br />

Toxicocladosporium irritans Crous & U. Braun, sp. nov.<br />

MycoBank MB504427. Fig. 4.<br />

Etymology: Named after the skin irritation resulting from exposure<br />

to the fungus.<br />

Mycelium (in PDA) ex hyphis ramosis, septatis, atro-brunneis, minute verruculosis,<br />

(2–)3–4 µm latis, ultimo crassitunicatis et crassiseptatis. Conidiophora solitaria,<br />

dimorphosa, macronemata et solitaria vel micronemata. Conidiophora macronemata<br />

ex hyphis modice brunneis lateraliter oriunda, erecta, subcylindrica, recta,<br />

geniculata-sinuosa vel irregulariter curvata, non ramosa vel ad apicem ramosa,<br />

2–7-septata, atro-brunnea, leviter verruculosa, crassitunicata, septa atro-brunnea,<br />

30–60 × 4–6 µm; conidiophora micronemata saepe non septata, raro 1–2-septata,<br />

erecta, doliiformes vel subcylindrica, apicem versus leviter attenuata, 10–30 × 2.5–<br />

4 µm. Cellulae conidiogenae integratae, terminales vel laterales, subcylindricae,<br />

apicem versus leviter attenuatae, 7–12 × 3–4 µm, sympodiales, cum 1–3 locis<br />

conidiogenibus, denticulatis, 1–1.5 µm latis, incrassatis, fuscatis-refractivis. Conidia<br />

catenulata vel rami-catenulata, modice vel atro-brunnea, crassitunicata, septis<br />

incrassatis, fuscatis, levia vel subtile verruculosa; ramoconidia (0–)1(–3)-septata,<br />

constricta, late ellipsoidea vel subcylindrica, 7–15 × 3–5 µm; conidia secundaria<br />

ellipsoidea vel ovoidea, pallide vel modice brunnea, 0(–1)-septata, (5–)6–8(–10) ×<br />

(3–)4(–5) µm; hila protuberantes, 1–1.5 µm lata, hila ramoconidiorum incrassata<br />

et fuscata-refractiva, vel hila conidiorum secundariorum 0.5–1 µm lata et<br />

subconspicua.<br />

Mycelium on PDA consisting of branched, septate, dark brown, finely<br />

verruculose, (2–)3–4 µm wide hyphae; walls <strong>and</strong> septa becoming<br />

thickened <strong>and</strong> darkened with age. Conidiophores solitary, dimorphic,<br />

macronematous <strong>and</strong> solitary, or micronematous, reduced to<br />

conidiogenous cells. Macronematous conidiophores subcylindrical,<br />

straight to geniculate-sinuous, or irregularly curved, unbranched<br />

or branched above, 2–7-septate, dark brown, finely verruculose,<br />

walls thick, septa dark brown, 30–60 × 4–6 µm; medium brown<br />

hyphae giving rise to lateral, erect branches that become swollen,<br />

dark brown, <strong>and</strong> develop into macronematous conidiophores<br />

with thick-walled <strong>and</strong> dark septa; micronematous conidiophores<br />

39


Crous et al.<br />

Fig. 4. Toxicocladosporium irritans (type material). A–B, F. Microconidiophores. C–E. Macroconidiophores. G–H. Ramoconidia <strong>and</strong> conidia. Scale bars = 10 µm.<br />

aseptate, reduced to conidiogenous cells (rarely 1–2-septate, i.e.,<br />

with 1–2 supporting cells), erect, doliiform to subcylindrical, with<br />

slight taper towards the apex, 10–30 × 2.5–4 µm. Conidiogenous<br />

cells integrated, terminal or lateral, subcylindrical with slight taper<br />

towards apex, 7–12 × 3–4 µm; proliferating sympodially with 1–3<br />

apical loci that can be slightly protruding <strong>and</strong> denticle-like, 1–1.5<br />

µm wide, thickened, darkened <strong>and</strong> refractive. Conidia catenulate<br />

in branched or unbranched chains, medium to dark brown, thickwalled,<br />

with dark, thick septa, smooth to finely verruculose;<br />

ramoconidia (0–)1(–3)-septate, prominently constricted at septa,<br />

broadly ellipsoid to subcylindrical, 7–15 × 3–5 µm; conidia ellipsoid<br />

to ovoid, younger apical conidia pale to medium brown, 0(–1)-<br />

septate, (5–)6–8(–10) × (3–)4(–5) µm; hila protruding, 1–1.5 µm<br />

wide, thickened, darkened <strong>and</strong> refractive in ramoconidia, but less<br />

obvious in young conidia, where hila are 0.5–1 µm wide.<br />

Cultural characteristics: Colonies on PDA erumpent, spreading,<br />

with dense aerial mycelium <strong>and</strong> smooth, even margins; surface<br />

olivaceous-black (centre), olivaceous-grey in outer region; reverse<br />

olivaceous-black. Colonies reaching 35 mm diam after 1 mo at 25<br />

°C in the dark; colonies fertile.<br />

40


<strong>Cladosporium</strong> <strong>and</strong> morphologically <strong>similar</strong> genera<br />

Specimen examined: Suriname, Paramaribo, isolated from mouldy paint, Feb.<br />

1958, M.B. Schol-Schwarz, holotype <strong>CBS</strong>-H 19892, culture ex-type <strong>CBS</strong> 185.58.<br />

Notes: Toxicocladosporium irritans produces ample amounts<br />

of volatile metabolites, which cause a skin rash within minutes<br />

of opening an inoculated dish for microscopic examination.<br />

Morphologically <strong>and</strong> phylogenetically it is very <strong>similar</strong> to<br />

<strong>Cladosporium</strong> s. str., <strong>and</strong> produces dimorphic conidiophores, which<br />

is also commonly observed in <strong>Cladosporium</strong>. It is distinct by having<br />

dark, thick-walled conidial <strong>and</strong> conidiophore septa, <strong>and</strong> lacking the<br />

typical coronate <strong>Cladosporium</strong> scar type (David 1997).<br />

Verrucocladosporium K. Schub., Aptroot & Crous, gen. nov.<br />

MycoBank MB504432.<br />

Etymology: Named after its frequently coarsely verrucose to<br />

warted hyphae, conidiophores <strong>and</strong> conidia, <strong>and</strong> cladosporium-like<br />

morphology.<br />

Differt a Cladosporio hyphis saepe verrucosis, hyalinis, conidiophoris cylindraceisfiliformibus,<br />

rectis, non vel vix geniculatis, non nodulosis, locis conidiogenis leviter<br />

incrassatis, distincte fuscatis-refractivis, sed non coronatis, conidiis saepe valde<br />

variantibus, saepe irregulariter formatis, grosse verrucosis-rugosis.<br />

Mycelium sparingly branched, hyphae septate, not constricted at<br />

septa, hyaline, almost smooth to irregularly rough-walled, coarsely<br />

verrucose to warted. Conidiophores arising laterally from creeping<br />

hyphae, erect, straight, or somewhat flexuous, narrowly cylindrical<br />

to filiform, neither geniculate nor nodulose, unbranched, septate,<br />

pale brown, thin-walled, smooth to often irregularly rough-walled or<br />

verrucose. Conidiogenous cells integrated, terminal or intercalary,<br />

cylindrical, polyblastic, with sympodial proliferation, with loci often<br />

crowded at the apex, truncate, barely to slightly thickened, but<br />

distinctly darkened-refractive. Ramoconidia cylindrical, aseptate,<br />

concolorous with conidiophores, thin-walled, irregularly roughwalled,<br />

coarsely verruculose to verrucose-rugose; hila unthickened<br />

but somewhat refractive. Conidia in long unbranched or loosely<br />

branched chains, obovoid, ellipsoid, fusiform to subcylindrical, with<br />

swollen <strong>and</strong> constricted parts, often appearing irregular in shape<br />

<strong>and</strong> outline, 0–1-septate, pale brown, thin-walled <strong>and</strong> irregularly<br />

rough-walled, verruculose-rugose; hila truncate, barely to slightly<br />

thickened, but distinctly darkened-refractive.<br />

Type species: Verrucocladosporium dirinae K. Schub., Aptroot &<br />

Crous, sp. nov.<br />

Verrucocladosporium dirinae K. Schub., Aptroot & Crous, sp.<br />

nov. MycoBank MB504433. Fig. 5.<br />

Etymology: Named after its host, Dirina massiliensis.<br />

Mycelium sparse ramosum. Hyphae 1–3 µm latae, septatae, non constrictae,<br />

hyalinae, leviae, vel irregulariter verruculosae, interdum verrucosae, tuberculatae,<br />

tenuitunicatae. Conidiophora ex hyphis repentibus lateraliter oriunda, erecta,<br />

recta, interdum leviter flexuosa, anguste cylindrica vel filiformes, non geniculta,<br />

non nodulosa, non ramosa, ad 85 µm longa, 2–3 µm lata, septata, tenuitunicata<br />

(≤ 0.75 µm), pallide brunnea, levia vel saepe irregulariter verrucosa, leviter<br />

crassitunicata. Cellulae conidiogenae integratae, saepe terminales, interdum<br />

intercalares, cylindricae, angustae, 9–20 µm longae, holoblasticae, sympodiales,<br />

locis conidiogenibus 1–3, saepe ad apicem aggregatis, interdum protuberantibus,<br />

truncatis, 1–1.8(–2) µm latis, incrassatis et fuscatis-refractivis. Ramoconidia<br />

cylindrica, 16–21 × (2–)2.5–3 µm, non septata, pallide brunnea, tenuitunicata,<br />

irregulariter verruculosa vel crosse verrucosa-rugosa, ad 4 hilis terminalibus,<br />

ad basim late truncata, non attenuata, 2–2.5 µm lata, non incrassata, sed leviter<br />

refractiva. Conidia catenata, in catenis longis, non ramosis vel laxe ramosis, plus<br />

minusve recta, obovoidea, ellipsoidea, fusiformes vel subcylindricae, sed saepe<br />

irregulares, 4–18(–23) × (2–)2.5–3.5 µm, 0–1-septata, ad septa interdum constricta,<br />

pallide brunnea, tenuitunicata (≤ 0.5 µm), irregulariter verruculosa-rugosa, utrinque<br />

leviter attenuata, hila truncata, (0.5–)0.8–1.5(–2) µm lata, vix vel leniter incrassata,<br />

sed distincte fuscata-refractiva.<br />

Mycelium sparingly branched; hyphae 1–3 µm wide, septate, not<br />

constricted at septa, hyaline, smooth to irregularly rough-walled,<br />

sometimes coarsely verrucose, with small to large drop-like,<br />

tuberculate warts, walls unthickened. Conidiophores arising laterally<br />

from creeping hyphae, erect, straight, sometimes slightly flexuous,<br />

narrowly cylindrical to filiform, not geniculate, non nodulose,<br />

unbranched, up to 85 µm long, 2–3 µm wide, septate, thin-walled<br />

(≤ 0.75 µm), pale brown, smooth to often irregularly rough-walled,<br />

verrucose, walls slightly thickened. Conidiogenous cells integrated,<br />

mostly terminal, sometimes also intercalary, cylindrical, narrow,<br />

9–20 µm long, conidiogenesis holoblastic, proliferation sympodial,<br />

with a single or up to three conidiogenous loci, often crowded at<br />

the apex, sometimes situated on small lateral prolongations, loci<br />

truncate, 1–1.8(–2) µm wide, thickened <strong>and</strong> darkened-refractive.<br />

Ramoconidia cylindrical, 16–21 × (2–)2.5–3 µm, aseptate,<br />

concolorous with conidiophores, thin-walled, irregularly roughwalled,<br />

verruculose to coarsely verrucose-rugose, apically with up<br />

to 4 hila, with a broadly truncate, non-attenuated base, 2–2.5 µm<br />

wide, unthickened but somewhat refractive. Conidia catenate, in<br />

long unbranched or loosely branched chains, more or less straight,<br />

obovoid, ellipsoid, fusiform to subcylindrical, but often appearing<br />

to form b<strong>and</strong>-like structures, with swollen <strong>and</strong> constricted parts,<br />

accordion or fir tree-like <strong>and</strong> also due to ornamentation often<br />

appearing irregular in shape <strong>and</strong> outline, 4–18(–23) × (2–)2.5–<br />

3.5 µm, 0–1-septate, sometimes constricted at the more or less<br />

median septum, pale brown, thin-walled (≤ 0.5 µm), irregularly<br />

rough-walled, verruculose-rugose, somewhat attenuated towards<br />

apex <strong>and</strong> base, hila truncate, (0.5–)0.8–1.5(–2) µm wide, barely<br />

or slightly thickened, but distinctly darkened-refractive; microcyclic<br />

conidiogenesis not observed.<br />

Cultural characteristics: Colonies erumpent, spreading, with<br />

catenate, feathery margins <strong>and</strong> moderate aerial mycelium on PDA.<br />

Surface grey-olivaceous, reverse iron-grey. Colonies reaching 25<br />

mm after 1 mo at 25 °C.<br />

Specimen examined: U.K., Somerset, Kingsbury Episcopi, isolated from the lichen<br />

Dirina massiliensis (Roccelaceae, Arthoniales), Mar. 2003, A. Aptroot, holotype<br />

<strong>CBS</strong>-H 19883, culture ex-type <strong>CBS</strong> 112794.<br />

Notes: Verrucocladosporium dirinae was deposited as<br />

<strong>Cladosporium</strong> arthoniae M. Christ. & D. Hawksw., but the name<br />

was misapplied. <strong>The</strong> latter species, described from apothecia of<br />

Arthonia impolita on Quercus from Sweden, does not possess<br />

clearly visible, distinct conidiogenous loci <strong>and</strong> hila, <strong>and</strong> therefore<br />

has to be excluded from the <strong>genus</strong> <strong>Cladosporium</strong> s. str. <strong>and</strong> is also<br />

easily distinguishable from the newly introduced species above.<br />

Furthermore the conidiophores are apically frequently branched <strong>and</strong><br />

the catenate, ellipsoid conidia are smaller <strong>and</strong> wider, 6–10 × 4–5 µm<br />

(Hawksworth 1979). Due to the conidiogenesis <strong>and</strong> the structure of<br />

the conidiogenous loci <strong>and</strong> conidia, C. arthoniae is rather close to<br />

lichenicolous Taeniolella S. Hughes species. <strong>The</strong> unique feature<br />

of the new <strong>genus</strong> Verrucocladosporium is its unusual conidial <strong>and</strong><br />

hyphal ornamentation. Furthermore, it differs from <strong>Cladosporium</strong><br />

s. str. in having cylindrical-filiform conidiophores, which are neither<br />

geniculate nor nodulose, quite distinct, thickened <strong>and</strong> darkened,<br />

but non-coronate conidiogenous loci <strong>and</strong> often irregularly shaped<br />

conidia. Phylogenetially, it is also distinct as a sister taxon to<br />

<strong>Cladosporium</strong> s. str. Concerning differences to other cladosporioid<br />

genera, see “key to the genera”. Verrucocladosporium dirinae has<br />

been isolated from the lichen species Dirina massiliensis, i.e., this<br />

species is probably lichenicolous, although its ecology is not quite<br />

clear. Fruiting of this species in vivo has not yet been observed. A<br />

second unnamed, taeniolella-like, lichenicolous hyphomycete was<br />

also present on the thallus of this lichen.<br />

www.studiesinmycology.org<br />

41


Crous et al.<br />

Fig. 5. Verrucocladosporium dirinae (type material). A. Colonies on MEA. B–C. Conidial chains. D–H. Ramoconidia <strong>and</strong> conidia. Scale bars = 10 µm.<br />

Capnodiales, Teratosphaeriaceae<br />

Devriesia americana Crous & Dugan, sp. nov. MycoBank<br />

MB504434. Fig. 6.<br />

Etymology: Named after the geographic location of its type strain,<br />

New York, U.S.A.<br />

Differt a D. shelburniensi conidiophoris brevioribus (ad 30 µm longis), leviter<br />

latioribus (2–3 µm), ramoconidiis saepe nullis et conidiis 0–1-septatis.<br />

Mycelium consisting of branched, septate, 1.5–3 µm wide hyphae,<br />

irregular in width, predominantly guttulate, smooth, forming hyphal<br />

str<strong>and</strong>s <strong>and</strong> hyphal coils; hyphae frequently forming dark brown,<br />

thick-walled, intercalary, muriformly septate chlamydospores on<br />

PDA in culture. Conidiophores subcylindrical, medium brown,<br />

straight to irregularly curved, up to 7-septate <strong>and</strong> 30 µm tall, 2–3<br />

µm wide, or reduced to conidiogenous cells. Conidiogenous cells<br />

terminal or lateral on hyphae, 5–12 × 2–3 µm, medium brown,<br />

smooth, guttulate, subcylindrical, mono- to polyblastic; scars<br />

somewhat darkened <strong>and</strong> thickened, but not refractive. Conidia<br />

medium brown, guttulate, smooth, in mostly unbranched chains,<br />

subcylindrical to narrowly ellipsoidal, tapering towards subtruncate<br />

ends, 0–1-septate, (7–)8–12(–16) × 2(–2.5) µm; hila darkened,<br />

somewhat thickened, not refractive, 1–1.5 µm wide.<br />

42


<strong>Cladosporium</strong> <strong>and</strong> morphologically <strong>similar</strong> genera<br />

Fig. 6. Devriesia americana (type material). A–B. Chlamydospore-like structures formed in culture. C–F. Conidiophores giving rise to conidial chains. G–H. Conidia. Scale bars<br />

= 10 µm<br />

Cultural characteristics: Colonies erumpent, with sparse aerial<br />

mycelium on PDA, <strong>and</strong> smooth, uneven, wide margins, submerged<br />

under the agar surface; greenish-black (surface); reverse<br />

olivaceous-black; on OA iron-grey (surface). Colonies reaching<br />

8–15 mm diam on PDA after 14 d at 25 °C in the dark; colonies<br />

fertile, but sporulation sparse.<br />

Specimen examined: U.S.A., New York, Long Isl<strong>and</strong>, isolated from air, F.M. Dugan,<br />

holotype <strong>CBS</strong>-H 19894, culture ex-type <strong>CBS</strong> 117726 = ATCC 96545 = CPC 5121.<br />

Notes: Until recently, this species was treated as part of the<br />

“Phaeoramularia” hachijoensis species complex (Braun et al.<br />

2003). <strong>The</strong> present strain has conidia that are smaller than those<br />

of “Phaeoramularia” hachijoensis, which has ramoconidia that are<br />

1–3-septate, up to 30 µm long, <strong>and</strong> conidia that are predominantly<br />

1-septate, 10–21 × 2–4 µm (Matsushima 1975). From the illustration<br />

provided by Matsushima, it appears that “Phaeoramularia”<br />

hachijoensis is indeed a species of Pseudocladosporium U. Braun,<br />

a finding which is in agreement with the name Pseudocladosporium<br />

hachijoense (Matsush.) U. Braun proposed by Braun (1998).<br />

Devriesia americana is both morphologically <strong>and</strong> phylogenetically<br />

more allied to Teratosphaeria Syd. & P. Syd. than Venturia<br />

Sacc. Based on its pigmented conidiophores <strong>and</strong> catenulate<br />

conidia, <strong>and</strong> scars that are somewhat darkened <strong>and</strong> thickened,<br />

<strong>and</strong> the formation of chlamydospores in culture, it is allocated to<br />

Devriesia Seifert & N.L. Nick. Species of the <strong>genus</strong> Devriesia are<br />

ecologically different, however (Seifert et al. 2004), being soil-borne<br />

www.studiesinmycology.org<br />

<strong>and</strong> thermotolerant. It is possible, therefore, that further collections<br />

of this fungus may eventually indicate that it needs to be placed in a<br />

distinct <strong>genus</strong> within the Teratosphaeriaceae. Devriesia americana<br />

is the second species of Devriesia with muriform chlamydospores,<br />

beside D. shelburniensis N.L. Nick. & Seifert, but the latter species<br />

is easily distinguishable by its long <strong>and</strong> narrow conidiophores (ca<br />

100–200 × 1.5–2.5 µm) <strong>and</strong> abundant ramoconidia, up to 25.5<br />

µm long, with 0–3 septa. Furthermore, D. shelburniensis is a<br />

thermotolerant soil-borne hyphomycete.<br />

Stenella araguata Syd., Ann. Mycol. 28(1/2): 205. 1930. Figs 7–8.<br />

≡ <strong>Cladosporium</strong> araguatum (Syd.) Arx, Genera of Fungi Sporulating in<br />

pure Culture, Edn 2 (Vaduz): 224. 1974.<br />

= <strong>Cladosporium</strong> castellanii Borelli & Marcano, Castellania 1: 154. 1973.<br />

Leaf spots hypophyllous, irregular to subcircular, up to 8 mm<br />

diam, indistinct, yellow to pale brown with indistinct margins on<br />

IMI 15728(a); on IMI 34905 (Fig. 7) lesions are amphigenous, <strong>and</strong><br />

fascicles <strong>and</strong> sporodochia are rare, with superficial mycelium being<br />

predominant. Mycelium consisting of internal <strong>and</strong> external, medium<br />

brown, septate, branched, verruculose, 3–4 µm wide hyphae.<br />

Caespituli fasciculate to sporodochial, hypophyllous, medium brown,<br />

up to 120 µm wide <strong>and</strong> 60 µm high. Conidiophores arising singly<br />

from superficial mycelium, or aggregated in loose to dense fascicles<br />

arising from the upper cells of a brown stroma up to 70 µm wide <strong>and</strong><br />

30 µm high; conidiophores medium brown, finely verruculose, 1–5-<br />

septate, subcylindrical, straight to geniculate-sinuous, unbranched<br />

43


Crous et al.<br />

Fig. 7. Stenella araguata (syntype material, IMI 34905). A. Leaf spot. B. Conidiophore, conidia <strong>and</strong> verruculose hypha on leaf surface. C–D. Conidiophore with terminal<br />

conidiogenous cells. E–G. Ramoconidia <strong>and</strong> conidia. Scale bars = 10 µm.<br />

or branched, 20–40 × 3–4 µm. Conidiogenous cells terminal or<br />

lateral, unbranched, medium brown, finely verruculose, tapering<br />

to slightly or flat-tipped loci, proliferating sympodially, 5–20 × 3–4<br />

µm; scars thickened, darkened <strong>and</strong> refractive. Conidia solitary<br />

or catenulate, in simple chains, medium brown, verruculose,<br />

subcylindrical to narrowly obclavate, apex obtuse, base bluntly<br />

rounded with truncate hilum, straight, 0–3-septate, (7–)13–20(–25)<br />

× 3(–3.5) µm; hila thickened, darkened, refractive, 1–1.5 µm wide.<br />

Description based on <strong>CBS</strong> 105.75 (Fig. 8): Mycelium consisting<br />

of branched, septate, verruculose, medium brown, 2–4 µm wide<br />

hyphae. Conidiomata brown, superficial, sporodochial, up to 200 µm<br />

diam Conidiophores solitary, erect, micro- to macronematous, 1–12-<br />

septate, subcylindrical, straight to geniculate-sinuous or irregularly<br />

curved, 10–70 × 3–4 µm; frequently swollen <strong>and</strong> constricted at<br />

septa, thick-walled, medium brown, verruculose. Conidiogenous<br />

cells terminal <strong>and</strong> intercalary, subcylindrical, straight, but frequently<br />

branched laterally, 6–20 × 3–4 µm, with 1–3 flat-tipped loci that<br />

can be subdenticulate, 1.5–2 µm wide, somewhat darkened <strong>and</strong><br />

thickened, not prominently refractive. Conidia medium brown, thickwalled,<br />

verruculose, septa becoming darkly pigmented, occurring in<br />

branched chains. Ramoconidia subcylindrical to narrowly ellipsoid,<br />

12–25 × 3.5–4(–5) µm, 1(–4)-septate. Conidia occurring in short<br />

chains (–8), subcylindrical to narrowly ellipsoid, 0–1(–3)-septate,<br />

(7–)10–15(–20) × (2–)3–3.5(–4) µm; hila somewhat thickened,<br />

darkened but not refractive, 1.5–2(–2.5) µm wide.<br />

Cultural characteristics: Colonies on OA erumpent, spreading, with<br />

moderate aerial mycelium <strong>and</strong> smooth, even margins; olivaceousgrey<br />

(surface); on PDA olivaceous-black (surface), margins<br />

feathery, uneven, with moderate aerial mycelium; reverse iron-grey.<br />

Colonies reaching 20 mm diam after 1 mo at 25°C in the dark on<br />

OA.<br />

Specimens examined: Venezuela, Aragua, La Victoria, on leaf spots of<br />

Pithecellobium lanceolatum (Mimosaceae), Jan. 1928, H. Sydow, lectotype of S.<br />

araguata (selected here!) IMI 15728(a); 3 Feb. 1928, syntype of S. araguata, IMI<br />

34905. Venezuela, isolated from man with tinea nigra, 1973, D. Borelli, holotype of<br />

C. castellanii, IMI 183818, culture ex-type <strong>CBS</strong> 105.75.<br />

Notes: Stenella araguata is a leaf spot pathogen of Pithecellobium<br />

in Venezuela, <strong>and</strong> represents the type species of the <strong>genus</strong> Stenella<br />

[Two collections were cited, viz. no. 407, ‘La Victoria’, <strong>and</strong> no. 370,<br />

‘inter La Victoria et Suata’, both without any date, <strong>and</strong> without<br />

any specific type indication. Thus, the two collections have to be<br />

considered syntypes. <strong>The</strong> two IMI collections with different dates<br />

are parts of the syntypes, of which IMI 15728(a) is proposed here<br />

to serve as lectotype]. Stenella araguata was incorrectly seen as<br />

a species of <strong>Cladosporium</strong> by von Arx (1974), which has recently<br />

been morphologically circumscribed (Braun et al. 2003, Schubert et<br />

al. 2007b – this volume), <strong>and</strong> is linked to Davidiella teleomorphs.<br />

In a study by McGinnis & Padhye (1978), <strong>Cladosporium</strong><br />

castellanii (tinea nigra of human in Venezuela) was shown to be<br />

synonymous to Stenella araguata (leaf spots of Pithecellobium<br />

lanceolatum in Venezuela). In the present study we re-examined<br />

the ex-type strain of C. castellanii (<strong>CBS</strong> 105.75), <strong>and</strong> found conidia<br />

to be 0–1(–3)-septate, (7–)10–15(–20) × (2–)3–3.5(–4) µm, while<br />

those of the type specimen of S. araguata were <strong>similar</strong>, namely<br />

0–3-septate, (7–)13–20(–25) × 3(–3.5) µm. Furthermore, both<br />

collections have verruculose hyphae, which is the primary feature<br />

distinguishing Stenella from Passalora Fr. (Crous & Braun 2003).<br />

44


<strong>Cladosporium</strong> <strong>and</strong> morphologically <strong>similar</strong> genera<br />

Fig. 8. Stenella araguata (<strong>CBS</strong> 105.75). A–B. Conidiophore fascicles on a pine needle <strong>and</strong> tap-water agar, respectively. C–D, G. Conidiophores giving rise to conidial chains.<br />

E–F, H–J. Conidial chains with ramoconidia <strong>and</strong> conidia. Scale bars = 10 µm.<br />

Stenella has always been used for anamorphs of Mycosphaerella<br />

(Crous et al. 2004, 2006c), <strong>and</strong> the fact that it belongs to<br />

Teratosphaeria (Teratosphaeriaceae), <strong>and</strong> not Mycosphaerella<br />

(Mycosphaerellaceae), raised the question of how to treat stenellalike<br />

anamorphs in Mycosphaerella. Due to insufficient availability of<br />

cultures (Crous et al. 2000, 2001), the status of Stenella was left<br />

unresolved (Crous & Braun 2003). Presently (Crous & Groenewald,<br />

unpubl. data), it is clear that the stenella-like morphology type is<br />

polyphyletic within the Mycosphaerellaceae, <strong>and</strong> paraphyletic<br />

within the Capnodiales. Several species are known that represent<br />

morphological transitions between Stenella <strong>and</strong> Passalora. It<br />

seems logical, therefore, that future studies should favour using<br />

Passalora to also accommodate Mycosphaerella anamorphs with<br />

superficial, verruculose hyphae, which have traditionally been<br />

placed in Stenella. This is in spite of the fact that there are other<br />

generic names available within the Mycosphaerellaceae for taxa<br />

with a stenella-like morphology (pigmented structures, darkened,<br />

thickened, refractive scars, <strong>and</strong> superficial, verruculose mycelium),<br />

namely Zasmidium Fr. (1849) (see Arzanlou et al. 2007 – this<br />

volume), <strong>and</strong> Verrucisporota D.E. Shaw & Alcorn (1993). Based<br />

on the phylogenetic position of the type species, Stenella s. str.<br />

is an anamorph of Teratosphaeria (Teratosphaeriaceae). Using<br />

the generic concept as employed in this volume of the Studies in<br />

Mycology, however, the anamorph <strong>genus</strong> is accepted as being<br />

poly- <strong>and</strong> paraphyletic within the order Capnodiales.<br />

www.studiesinmycology.org<br />

45


Crous et al.<br />

Helotiales, incertae sedis<br />

Hyalodendriella Crous, gen. nov. MycoBank MB504435.<br />

Etymology: Morphologically <strong>similar</strong> to Hyalodendron Diddens.<br />

Differt a Hyalodendro et Retroconi conidiophoris dimorphis, cicatricibus incrassatis<br />

et conidiis ultimo brunneis.<br />

Morphologically <strong>similar</strong> to Hyalodendron <strong>and</strong> Retroconis, but<br />

distinct in that it has dimorphic conidiophores, conidia that turn<br />

brown with age, <strong>and</strong> have thickened scars. Microconidiophores<br />

forming as lateral branches on hyphae, subcylindrical, subhyaline<br />

to pale brown, smooth, septate, with terminal conidiogenous cells.<br />

Macroconidiophores septate, subcylindrical, straight to curved,<br />

subhyaline to pale brown, smooth, with an apical rachis that is pale<br />

brown, smooth, subcylindrical, with numerous, aggregated loci.<br />

Conidia limoniform to ellipsoid, aseptate, smooth, pale brown, in<br />

short chains, tapering towards ends that are prominently apiculate,<br />

prominently thickened <strong>and</strong> darkened, but not refractive.<br />

Type species: Hyalodendriella betulae Crous, sp. nov.<br />

Hyalodendriella betulae Crous sp. nov. MycoBank MB504436.<br />

Fig. 9.<br />

Mycelium ex hyphis ramosis, septatis, 1.5–2 µm latis, levibus, hyalinis vel pallide<br />

brunnei compositum. Conidiophora dimorphosa: (A) Conidiophora ex hyphis<br />

lateraliter oriunda, subcylindrical, subhyalina vel pallide brunnea, levia, 1–6-<br />

septata, ad 40 µm longa et 2–3 µm lata. Cellulae conidiogenae terminales, 5–15<br />

× 2–3 µm, loco conidiogeno singulare et terminale, cellula ellipsoidea (conidio ?),<br />

persistente, interdum cellulis catenulatis (ad 6), pallide brunneo, apice subacute<br />

rotundato, basi truncata, 5–7 × 3–4 µm. (B) Conidiophoris 10–20 × 2–3 µm,<br />

1–2-septatis, subcylindraceis, rectis vel curvatis, subhyalinis vel pallide brunneis,<br />

levibus. Cellulae conidiogenae pallide brunneae, leviae, subcylindraceae, locis<br />

numerosis, aggregatis, inconspicuis vel subdenticulatis, leviter protuberantes, 0.5<br />

µm diam, incrassatis et fuscatis. Conidia catenulata (2–3), (4–)5–6(–7) × 2.5–3 µm,<br />

limoniformes vel ellipsoidea, non septata, levia, pallide brunnea, utrinque attenuata,<br />

apiculata, 0.5–1 × 0.5 µm, incrassata et fuscata, non refractiva.<br />

Mycelium consisting of branched, septate, 1.5–2 µm wide<br />

hyphae, smooth, hyaline to pale brown. Conidiophores dimorphic.<br />

Type A: Conidiophores forming as lateral branches on hyphae,<br />

subcylindrical, subhyaline to pale brown, smooth, 1–6-septate, up<br />

to 40 µm long, <strong>and</strong> 2–3 µm wide. Conidiogenous cells terminal, 5–<br />

15 × 2–3 µm, with a single, apical locus, giving rise to an ellipsoidal<br />

cell (conidium?) which mostly remains attached, pale brown, with<br />

a subacutely rounded apex <strong>and</strong> truncate base, 5–7 × 3–4 µm, at<br />

times forming chains of up to 6 such cells. Type B: Conidiophores<br />

10–20 × 2–3 µm, 1–2-septate, subcylindrical, straight to curved,<br />

subhyaline to pale brown, smooth. Conidiogenous cells pale<br />

brown, smooth, subcylindrical with numerous, aggregated loci,<br />

inconspicuous to subdenticulate <strong>and</strong> somewhat protruding, 0.5 µm<br />

wide, somewhat thickened <strong>and</strong> darkened. Conidia in chains of 2–3,<br />

limoniform to ellipsoid, widest in the middle, aseptate, smooth, pale<br />

brown, tapering towards ends that are prominently apiculate, 0.5–1<br />

µm long, 0.5 µm wide, prominently thickened <strong>and</strong> darkened, but<br />

not refractive.<br />

Cultural characteristics: Colonies on PDA slimy, spreading,<br />

somewhat erumpent in the centre, with even, catenulate margins,<br />

lacking aerial mycelium; surface fuscous-black to olivaceous-black,<br />

with patches of cream; reverse fuscous-black with patches of<br />

cream. Colonies reaching 25 mm diam on PDA after 1 mo at 25 °C<br />

in the dark; colonies fertile with profuse sporulation on SNA.<br />

Specimen examined: Netherl<strong>and</strong>s, Oostelijk Flevol<strong>and</strong>, Jagersveld, isolated from<br />

Alnus glutinosa (Betulaceae), May 1982, W. Gams, holotype <strong>CBS</strong>-H 19895, culture<br />

ex-type <strong>CBS</strong> 261.82.<br />

Notes: Morphologically Hyalodendriella resembles the genera<br />

Hyalodendron <strong>and</strong> Retroconis de Hoog & Bat. Vegte (de Hoog<br />

& Batenburg van der Vegte 1989). It is distinct, however, in its<br />

pigmentation, dimorphic conidiophores <strong>and</strong> conidia. Furthermore,<br />

a strain of Retroconis fusiformis (S.M. Reddy & Bilgrami) de Hoog<br />

& Bat. Vegte (<strong>CBS</strong> 330.81) clusters apart from Hyalodendriella,<br />

namely in the Chaetomiaceae, Sordariales.<br />

Pleosporales, incertae sedis<br />

Ochrocladosporium Crous & U. Braun, gen. nov. MycoBank<br />

MB504437.<br />

Etymology: Named after its pale brown, cladosporium-like conidia.<br />

Differt a Cladosporio et generis cladosporioidibus diversis conidiophoris cum cellulis<br />

basalibus T-formibus et/vel cicatricibus non incrassatis, non vel leviter fuscatisrefractivis.<br />

Mycelium consisting of branched, septate hyphae, subhyaline to<br />

pale brown, smooth, giving rise to two types of conidiophores.<br />

Macronematous conidiophores solitary, erect, arising from<br />

superficial hyphae, composed of a subcylindrical stipe, without a<br />

swollen or lobed base or rhizoids, with or without a T-shaped foot cell,<br />

pale to dark brown; apical conidiogenous apparatus with or without<br />

additional branches, branched part, if present, with short branchlets<br />

composed of conidiogenous cells <strong>and</strong> ramoconidia, continuous to<br />

septate, wall thin or slightly thicked, pale brown. Conidiogenous cells<br />

integrated, terminal or intercalary, subcylindrical to doliiform, pale<br />

brown, thin-walled, smooth; unilocal or multilocal, determinate to<br />

sympodial, loci conically truncate, subdenticulate, neither thickened,<br />

nor darkened-refractive or only slightly darkened-refractive.<br />

Micronematous conidiophores integrated in hyphae, reduced to a<br />

lateral peg-like locus or erect, frequently reduced to conidiogenous<br />

cells, pale brown, smooth, subcylindrical. Conidia occurring in<br />

branched chains, fusiform, ellipsoid-ovoid to subcylindrical, 0(–1)-<br />

septate, ramoconidia present, pale brown, thin-walled, smooth to<br />

finely verruculose, ends attenuated, hila obconically truncate to<br />

almost pointed, neither thickened nor darkened-refractive.<br />

Type species: Ochrocladosporium elatum (Harz) Crous & U. Braun,<br />

comb. nov.<br />

Ochrocladosporium elatum (Harz) Crous & U. Braun, comb.<br />

nov. MycoBank MB504438. Fig. 10.<br />

Basionym: Hormodendrum elatum Harz, Bull. Soc. Imp. Naturalistes<br />

Moscou 44: 140. 1871.<br />

≡ <strong>Cladosporium</strong> elatum (Harz) Nannf., in Melin & Nannfeldt, Svenska<br />

Skogsvardsfoereren Tidskr. 32: 397. 1934.<br />

≡ Cadophora elatum (Harz) Nannf., in Melin & Nannfeldt, Svenska<br />

Skogsvardsfoereren Tidskr. 32: 422. 1934.<br />

Mycelium consisting of branched, septate, smooth, hyaline, 2–4<br />

µm wide, thin-walled, hyphae, becoming darker brown in places,<br />

giving rise to erect conidiophores. Conidiophores either reduced<br />

to conidiogenous cells, or well-differentiated, terminal <strong>and</strong> lateral<br />

on hyphae, erect, highly variable, arising from superficial <strong>and</strong><br />

submerged hyphae, reduced to subdenticulate loci, 1–1.5 µm<br />

wide, or well-differentiated, up to 60 µm long, 1–3-septate, 3–4<br />

µm wide, hyaline to medium brown, smooth, thin-walled (≤ 1 µm).<br />

Conidiogenous cells integrated as lateral peg-like loci on hyphal<br />

cells, or erect, subcylindrical, up to 25 µm long, 2.5–4 µm wide,<br />

with 1–3 terminal loci, occasionally also lateral, 1–1.5 µm wide,<br />

not thickened <strong>and</strong> darkened, but frequently somewhat refractive<br />

(mounted in Shear’s solution, not lactic acid). Ramoconidia<br />

46


<strong>Cladosporium</strong> <strong>and</strong> morphologically <strong>similar</strong> genera<br />

Fig. 9. Hyalodendriella betulae (type material). A. Conidiophores on PDA. B–C. Microconidiophores. D–H. Macroconidiophores with fascicles of conidiogenous cells. I. Conidia<br />

with darkened, thickened hila. Scale bars = 10 µm.<br />

subcylindrical to ellipsoid, hyaline to pale brown, smooth to finely<br />

verruculose, 10–40 × 3–5 µm, 0(–1)-septate, giving rise to branched<br />

chains of conidia (up to 20 per chain) that are subcylindrical to<br />

ellipsoid, aseptate, (7–)8–10(–14) × (3–)4(–4.5) µm, smooth to<br />

finely verruculose, olivaceous-brown, thin-walled (up to 0.5 µm),<br />

hila 0.5–1 µm wide, neither thickened nor, or barely, darkened<br />

refractive.<br />

Cultural characteristics: Colonies erumpent, spreading, fast<br />

growing, covering the plate within 1 mo at 25 °C; aerial mycelium<br />

abundant, margins smooth on PDA; surface isabelline in centre,<br />

umber in outer region; olivaceous-black in reverse.<br />

Specimen examined: Sweden, Iggesund, isolated from wood pulp, Jan. 1976, E.<br />

Melin, specimen <strong>CBS</strong>-H 19896, culture <strong>CBS</strong> 146.33.<br />

Notes: “Hormodendrum” elatum was originally described from<br />

a wooden stump in Germany. <strong>The</strong> culture examined here was<br />

deposited by Melin in 1933 as culture 389:14, isolated from wood<br />

chips in Sweden, <strong>and</strong> described by Nannfeldt, <strong>and</strong> has since<br />

been accepted as authentic for the species. Earlier publications<br />

(de Vries 1952, Ho et al. 1999, de Hoog et al. 2000), clearly state<br />

that this species does not belong in <strong>Cladosporium</strong> s. str., <strong>and</strong> this<br />

statement is supported by the phylogenetic analysis placing it in<br />

the Pleosporales.<br />

www.studiesinmycology.org<br />

Ochrocladosporium frigidarii Crous & U. Braun, sp. nov.<br />

MycoBank MB504439. Fig. 11.<br />

Etymology: Named after it collection site, within a cooled incubation<br />

room.<br />

Differt a O. elato conidiophoris distincte dimorphis, macroconidiophoris majoribus,<br />

ad 600 × 5–7 µm, septis incrassates, cellulis basalibus T-formibus et conidiis leniter<br />

brevioribus et latioribus, (6–)7–8(–10) × (4–)4.5–5(–6) µm.<br />

Mycelium consisting of branched, septate, 2–7 µm wide hyphae,<br />

occasionally constricted at septa with hyphal swellings, subhyaline<br />

to pale brown, smooth, thin-walled, giving rise to two types of<br />

conidiophores. Macronematous conidiophores solitary, erect,<br />

arising from superficial hyphae, up to 600 µm long, composed of<br />

a subcylindrical stipe, 5–7 µm wide, 10–15(–20)-septate, without<br />

a swollen or lobed base or rhizoids, but with a T-shaped foot cell,<br />

wall ≤ 1 µm wide, guttulate, with thick septa, dark brown, finely<br />

verruculose, apical 1–2 cells at times medium brown, giving rise<br />

to 1–2 primary branches, 0–1-septate, subcylindrical, thin-walled,<br />

pale brown, smooth to finely verruculose, 10–20 × 4–6 µm, giving<br />

rise to (1–)2–4 secondary branches, 0–1-septate, subcylindrical,<br />

8–13(–20) × 4–5 µm, or giving rise directly to conidiogenous cells.<br />

Conidiogenous cells subcylindrical to doliiform, pale brown, smooth,<br />

8–15 × 3–4 µm, loci somewhat protruding 1–2 µm wide, neither<br />

47


Crous et al.<br />

Fig. 10. Ochrocladosporium elatum (<strong>CBS</strong> 146.33). A–C, E. Microconidiophores. D. Macro- <strong>and</strong> microconidiophore. F–H. Macroconidiophores. I–J. Ramoconidia <strong>and</strong> conidia.<br />

Scale bars = 10 µm.<br />

thickened, darkened, nor refractive. Micronematous conidiophores<br />

erect, pale brown, smooth, subcylindrical, reduced to conidiogenous<br />

cells, or up to 4-septate, 15–90 × 2–3.5 µm, mostly unbranched,<br />

rarely branched below; conidiogenous cells subcylindrical,<br />

pale brown, smooth to finely verruculose, tapering at apex <strong>and</strong><br />

sometimes at base, proliferating sympodially via 1(–3) loci, 1–1.5<br />

µm wide, denticle-like, which can appear somewhat darkened;<br />

micronematous conidiophores frequently occurring at the base of<br />

macronematous conidiophores. Ramoconidia, if present, up to 30<br />

µm long, 0–1-septate. Conidia <strong>and</strong> ramoconidia ellipsoid to ovoid,<br />

aseptate, pale brown, thin-walled (≤ 0.75 µm), finely verruculose,<br />

occurring in branched chains; conidia (6–)7–8(–10) × (4–)4.5–5(–6)<br />

µm; hila 0.5–1 µm wide, not darkened, thickened or refractive.<br />

Cultural characteristics: Colonies on PDA erumpent, spreading, with<br />

profuse sporulation <strong>and</strong> moderate aerial mycelium, even margins,<br />

olivaceous-grey (surface); reverse olivaceous-black. Colonies<br />

covering the dish after 1 mo at 25 °C in the dark.<br />

Specimen examined: Germany, Hannover, isolated from a cooled room, Jan. 1981,<br />

B. Ahlert, holotype <strong>CBS</strong>-H 19897, culture ex-type <strong>CBS</strong> 103.81.<br />

Notes: Ochrocladosporium frigidarii is characterised by its<br />

dimorphic fruiting, <strong>and</strong> inconspicuous scars <strong>and</strong> conidial hila, which<br />

48


<strong>Cladosporium</strong> <strong>and</strong> morphologically <strong>similar</strong> genera<br />

Fig. 11. Ochrocladosporium frigidarii (type material). A. Macro- <strong>and</strong> microconidiophores. B. Foot cell of macroconidiophore. C–D. Microconidiophore. E–J. Macroconidiophores.<br />

K. Conidia. Scale bars = 10 µm.<br />

are distinct from <strong>Cladosporium</strong> s. str. <strong>The</strong> phylogenetic analysis of<br />

its LSU sequence places it in the Pleosporales, together with O.<br />

elatum.<br />

<strong>The</strong> dimorphic conidiophores seen in O. frigidarii (<strong>CBS</strong> 103.81)<br />

are less obvious in O. elatum (<strong>CBS</strong> 146.33), but the scars <strong>and</strong> hila<br />

are <strong>similar</strong>. <strong>The</strong> macronematous conidiophores of O. frigidarii are<br />

much longer <strong>and</strong> wider <strong>and</strong> the conidia are shorter <strong>and</strong> slightly<br />

wider, (6–)7–8(–10) × (4–)4.5–5(–6) µm, than those of O. elatum<br />

which are (7–)8–10(–14) × (3–)4(–4.5) µm.<br />

www.studiesinmycology.org<br />

49


Crous et al.<br />

Fig. 12. Rhizocladosporium argillaceum (type material). A–E. Conidiophores with pigmented ramoconidia <strong>and</strong> hyaline conidia. F. Rhizoids forming at the foot cells of<br />

macroconidiophores. Scale bar = 10 µm.<br />

Incertae sedis<br />

Rhizocladosporium Crous & U. Braun, gen. nov. MycoBank<br />

MB504440.<br />

Etymology: Named after the presence of rhizoids on its<br />

conidiophores, <strong>and</strong> cladosporium-like conidia.<br />

Differt a Cladosporio et generis cladosporioidibus diversis hyphis hyalinis,<br />

conidiophoris cum cellulis basalibus lobatis vel rhizoidibus, cellulis conidiogenis<br />

monoblasticis, determinatis, locis margine leviter incrassatis et fuscatis, non<br />

refractivis, non coronatis, ramoconidiis brunneis sed conidiis hyalinis, hilis non<br />

incrassatis, non fuscatis-refractivis.<br />

Mycelium consisting of branched, septate, smooth, hyaline hyphae.<br />

Conidiophores solitary, macronematous, subcylindrical, erect,<br />

arising from superficial mycelium, septate, pigmented, smooth;<br />

base somewhat inflated, lobed or with rhizoids. Conidiogenous<br />

cells integrated, terminal, monoblastic, determinate, subcylindrical,<br />

tapering towards a single flat-tipped locus, straight to once<br />

geniculate, occasionally with two loci, pigmented, smooth; locus<br />

flattened, undifferentiated to somewhat darkened <strong>and</strong> thickened<br />

along the rim, not refractive, giving rise to a single conidial chain<br />

or a single ramoconidium with several simple acropetal chains of<br />

secondary ramoconidia or conidia. Conidia occurring in branched<br />

chains; ramoconidia subcylindrical to narrowly ellipsoidal, straight<br />

to geniculate-sinuous, with apical <strong>and</strong> lateral conidial hila;<br />

ramoconidia with broadly truncate base medium brown; secondary<br />

ramoconidia with narrowed base subhyaline or hyaline, smooth;<br />

conidia aseptate, in chains, hyaline, guttulate, ellipsoidal with<br />

obtuse ends; hila inconspicuous, neither darkened nor refractive<br />

or thickened.<br />

Type species: Rhizocladosporium argillaceum (Minoura) Crous &<br />

U. Braun, comb. nov.<br />

Rhizocladosporium argillaceum (Minoura) Crous & U. Braun,<br />

comb. nov. MycoBank MB504441. Fig. 12.<br />

Basionym: <strong>Cladosporium</strong> argillaceum Minoura, J. Ferment. Technol.<br />

44: 140. 1966.<br />

Mycelium consisting of branched, septate, smooth, hyaline,<br />

thin-walled, 1.5–2 µm wide hyphae. Conidiophores solitary,<br />

macronematous, erect, arising from superficial mycelium; base<br />

somewhat inflated, lobed or with rhizoids, up to 10 µm wide;<br />

conidiophore stipe subcylindrical, straight to curved, rarely<br />

geniculate-sinuous, wall up to 1 µm wide, medium brown,<br />

sometimes paler towards the tip, smooth, 1–6-septate, 35–160<br />

50


<strong>Cladosporium</strong> <strong>and</strong> morphologically <strong>similar</strong> genera<br />

µm tall, 4–6 µm wide. Conidiogenous cells terminal, straight,<br />

subcylindrical, tapering towards a flat-tipped locus, occasionally<br />

once geniculate, with two loci, medium brown, smooth, 15–35 ×<br />

4–6 µm; locus flattened, undifferentiated or very slightly darkened<br />

<strong>and</strong> thickened along the rim, not refractive, 1.5–2 µm wide. Conidia<br />

occurring in branched chains. Ramoconidia subcylindrical to<br />

narrowly ellipsoidal, straight to geniculate-sinuous, 17–35 × 4–5<br />

µm, medium brown, smooth, thin-walled, frequently branching<br />

laterally, with apical <strong>and</strong> lateral subdenticulate conidial hila, 1.5–2.5<br />

µm wide; secondary ramoconidia hyaline or subhyaline. Conidia<br />

aseptate, (10–)12–17(–20) × (3.5–)4(–4.5) µm, in branched<br />

chains (–6), hyaline or subhyaline, guttulate, ellipsoidal-fusiform,<br />

with obtuse ends, or tapering to obconically subtruncate ends with<br />

hila that are inconspicuous (neither darkened nor refractive or<br />

thickened), 0.5–1 µm wide.<br />

Cultural characteristics: Colonies on PDA spreading, erumpent, with<br />

smooth, even margins <strong>and</strong> sparse to moderate aerial mycelium;<br />

hazel to fawn (surface); reverse hazel to fawn. Colonies reaching<br />

35 mm diam after 1 mo at 25 °C in the dark; colonies fertile.<br />

Specimen examined: Japan, Yoku Isl<strong>and</strong>, isolated from decayed myxomycete, 21<br />

Oct. 1961, K. Tubaki No. 4262 holotype, culture ex-type <strong>CBS</strong> 241.67 = IFO 7055.<br />

Notes: <strong>The</strong> lobed-rhizoid conidiophore base, <strong>and</strong> brown,<br />

disarticulating ramoconidia, with hyaline chains of conidia, are<br />

characteristic of Rhizocladosporium. Although Minoura (1966)<br />

illustrated some conidiophores that were micronematous (reduced<br />

to conidiogenous cells on superficial mycelium), these were<br />

not observed in the present study. Metulocladosporiella Crous,<br />

Schroers, J.Z. Groenew., U. Braun & K. Schub. (Crous et al.<br />

2006a) (Herpotrichiellaceae), comprising two banana leaf-spotting<br />

pathogens, is another cladosporioid hyphomycete <strong>genus</strong> having<br />

distinct rhizoid hyphae at the swollen base of conidiophores. It<br />

differs, however, in having conidiophores terminally branched in<br />

a metula-like manner <strong>and</strong> distinct conidiogenous loci <strong>and</strong> conidial<br />

hila. Pleurotheciopsis B. Sutton (Ellis 1976) is also characterised<br />

by having pigmented conidiophores <strong>and</strong> hyaline or pale, septate<br />

conidia formed in acropetal chains, but the conidiophores<br />

proliferate percurrently, the conidiogenous cells are polyblastic <strong>and</strong><br />

ramoconidia are lacking, i.e., the conidia are formed in unbranched<br />

chains. Parapleurotheciopsis P.M. Kirk (Kirk 1982) is very <strong>similar</strong> to<br />

Rhizocladosporium. <strong>The</strong> conidiophores possess a single terminal<br />

unilocal conidiogenous cell giving rise to a single ramoconidium<br />

which forms several chains of acropetal, aseptate, hyaline to pale<br />

olivaceous conidia. <strong>The</strong> base of the conidiophores is somewhat<br />

swollen <strong>and</strong> lobed [except for Parapleurotheciopsis coccolobae<br />

R.F. Castañeda & B. Kendr., Castañeda & Kendrick (1990), with at<br />

most slightly swollen, but unlobed base]. However, R. argillaceum<br />

occasionally has once-geniculate conidiogenous cells with two<br />

loci. Furthermore, it clusters in the Helotiales (Fig. 1), whereas a<br />

sequenced strain of Parapleurotheciopsis inaequiseptata (MUCL<br />

41089), belongs to the Xylariales (Fig. 2). <strong>The</strong> occasionally<br />

occurring conidiogenous cells with two loci <strong>and</strong> the aseptate conidia<br />

connect Rhizocladosporium with Subramaniomyces Varghese &<br />

V.G. Rao (Varghese & Rao 1979, Kirk 1982) in which, however,<br />

terminal ramoconidia are lacking. Furthermore, the type species,<br />

S. fusisaprophyticus (Matsush.) P.M. Kirk, frequently has branched<br />

conidiophores. Subramaniomyces simplex U. Braun & C.F. Hill<br />

(Braun & Hill 2002), a species with unbranched conidiophores is,<br />

however, morphologically <strong>similar</strong> to R. argillaceum, but the <strong>genus</strong><br />

Subramaniomyces is phylogenetically distinct <strong>and</strong> also belongs to<br />

the Xylariales (<strong>CBS</strong> 418.95, Fig. 2).<br />

Key to <strong>Cladosporium</strong> <strong>and</strong> morphologically <strong>similar</strong> genera<br />

(bearing simple or branched acropetal chains of amero- to phragmosporous blastoconidia)<br />

1. Conidiophores <strong>and</strong> conidia hyaline............................................................................................................................................................. 2<br />

1. At least conidiophores pigmented .............................................................................................................................................................. 4<br />

2. Conidia in simple chains ........................................................................................................................................................... Hormiactis<br />

2. Conidia in branched chains ....................................................................................................................................................................... 3<br />

3. Conidiogenous cells sympodial, with distinct conidiogenous loci (scars), thickened <strong>and</strong> darkened; conidia amero- to phragmosporous;<br />

plant pathogenic, leaf-spotting fungi (Mycosphaerella anamorphs; Mycosphaerellaceae) ......................................................... Ramularia<br />

3. Terminal conidiogenous cells with denticle-like loci, giving rise to ramoconidia which form simple or branched conidial chains;<br />

lignicolous ........................................................................................................................................................................... Hyalodendron<br />

[Conidiophores dimorphic; mycelium, conidiophores <strong>and</strong> conidia at first hyaline, later turning pale brown; conidia in short chains, see<br />

Hyalodendriella]<br />

4(1). Conidia distoseptate, in simple chains .............................................................................................................................................. Lylea<br />

4. Conidia aseptate or euseptate ................................................................................................................................................................... 5<br />

5. Conidiophores little differentiated, micronematous to semimacronematous; conidiogenous loci undifferentiated, truncate, neither<br />

distinctly thickened nor darkened or only very slightly so .......................................................................................................................... 6<br />

5. Conidiophores well-differentiated, semimacronematous (but multilocal <strong>and</strong>/or conidiogenous loci well-differentiated) to<br />

macronematous ....................................................................................................................................................................................... 12<br />

6. Conidiophores <strong>and</strong> conidia delicate, thin-walled, in long, easily disarticulating chains .............................................................................. 7<br />

6. Conidiophores <strong>and</strong> conidia robust, wall thickened, dark, conidial chains often seceding with difficulty .................................................... 9<br />

7. Conidiophores semimacronematous, simple to often irregularly branched; conidia delicate, narrow, 1–3 µm wide, hyaline to pale<br />

olivaceous ............................................................................................................................................................................ Polyscytalum<br />

www.studiesinmycology.org<br />

51


Crous et al.<br />

7. Conidiophores unbranched, micronematous or semimicronematous, integrated in ordinary hyphae, forming minute, lateral,<br />

monoblastic, determinate, peg-like protuberances to semimacronematous, forming short lateral branches (conidiophores) with several<br />

inconspicuous to denticle-like loci .............................................................................................................................................................. 8<br />

8. Phialidic synanamorphs often present, but sometimes also lacking; saprobic, rarely plant pathogenic, often human pathogenic<br />

(Herpotrichiellaceae, Chaetothyriales) .......................................................................................................................... Cladophialophora<br />

8. Without phialidic synanamorphs; saprobic or plant pathogenic (Venturia, Venturiaceae)..............................................................................<br />

.......................................................................................................................................... Fusicladium s. lat. (incl. Pseudocladosporium)<br />

9(6). Conidia aseptate, rarely 1-septate; lignicolous, on dead wood ............................................................................................... Xylohypha<br />

9. Conidia septate ........................................................................................................................................................................................ 10<br />

10. Conidia 1-septate, with a dark brown to blackish b<strong>and</strong> at the septum; on dead wood .................................................................. Bispora<br />

10. Conidia at least partly 2- to pluriseptate <strong>and</strong>/or without dark brown to blackish b<strong>and</strong> at the septum ...................................................... 11<br />

11. Conidia branched ....................................................................................................................................................................... Taeniolina<br />

11. Conidia unbranched ................................................................................................................................................................... Taeniolella<br />

12(5). Conidiogenous loci <strong>and</strong> conidial hila distinctly coronate, i.e., composed of a central convex dome surrounded by a periclinal raised rim,<br />

mostly at least somewhat protuberant (anamorphs of Davidiella, Davidiellaceae, Capnodiales) ............................. <strong>Cladosporium</strong> s. str.<br />

12. Conidiogenous loci non-coronate (either inconspicuous, thickened <strong>and</strong> darkened or denticle-like) ........................................................ 13<br />

13. Mycelium, conidiophores <strong>and</strong> conidia at first hyaline or subhyaline, later turning pale brown; conidiophores dimorphic, either<br />

conidiogenous cells with a single conidiogenous locus, giving rise to an ellipsoid cell (conidium?) which mostly remains attached,<br />

base truncate, apex subacutely rounded, at times forming chains of such cells; or conidiophores with numerous aggregated loci,<br />

inconspicuous to subdenticulate; conidia in short chains, of mostly 2–3 (isolated from Alnus in Europe) ........................ Hyalodendriella<br />

13. Fruiting different; at least conidiophores consistently pigmented or conidiophores uniform or loci distinct; conidia mostly in long, often<br />

branched chains ...................................................................................................................................................................................... 14<br />

14. Conidiophores with verruculose conidiogenous apices, otherwise smooth; conidia distinctly verruculose-verrucose; conidiogenous loci<br />

<strong>and</strong> conidial hila inconspicuous ............................................................................................................................................................... 15<br />

14. Conidiophores either smooth throughout or verruculose below <strong>and</strong> smooth above or verruculose throughout; <strong>and</strong>/or conidiogenous loci<br />

conspicuous, i.e., thickened <strong>and</strong> darkened or denticle-like ...................................................................................................................... 16<br />

15. Conidiophores macronematous, unbranched, base swollen, with percurrent regenerative proliferations, unrelated to conidiation;<br />

conidiogenous cells terminal, occasionally also subterminal; conidia terminally <strong>and</strong> laterally formed, aseptate (saprobic on leaves)<br />

................................................................................................................................................................................................ Castanedaea<br />

15. Conidiophores little differentiated, semimacronematous, unbranched or with short lateral branchlets, base undifferentiated,<br />

without percurrent proliferations; conidiogenous cells terminal <strong>and</strong> occasionally intercalary-pleurogenous; conidia terminally <strong>and</strong><br />

subterminally formed, 0–2-septate (lignicolous, on decorticated wood) .......................................................................... Websteromyces<br />

16(14). Conidiophores unbranched, with a simple terminal conidiogenous cell, non-geniculate-sinuous, subcylindrical to somewhat inflated at<br />

the tip; conidiogenous loci terminal <strong>and</strong> lateral, inconspicuous or subconspicuous, neither thickened nor darkened, non-protuberant;<br />

conidia attached with a very narrow, pointed hilum ................................................................................................................................. 17<br />

16. Conidiophores with a branched terminal conidiogenous apparatus, composed of conidiogenous cells <strong>and</strong>/or ramoconidia or conidiophores<br />

unbranched, with a single terminal conidiogenous cell or additional intercalary ones, but conidiogenous loci different, conspicuous,<br />

thickened <strong>and</strong> darkened or denticle-like .................................................................................................................................................. 18<br />

17. Conidiophores with distinct rhizoid-digitate base; tips of the conidiogenous cells somewhat swollen, usually unilaterally swollen or<br />

somewhat curved; conidia solitary or only in very short unbranched chains; hyperparasitic on rusts .................................. Digitopodium<br />

17. Conidiophores without rhizoid-digitate base; tips of the conidiogenous cells subcylindrical to somewhat swollen, but swellings not unilateral<br />

<strong>and</strong> not curved; conidia solitary <strong>and</strong> in simple or branched chains; associated with leaf spots ................................. Rachicladosporium<br />

18(16). Conidiophores in synnematous conidiomata ..................................................................................................................................... 19<br />

18. Synnemata lacking .................................................................................................................................................................................. 20<br />

19. Conidiogenous cells with a single or several truncate to subdenticulate, relatively broad conidiogenous loci; conidia with truncate, flat hila;<br />

on wood, resin ............................................................................................................................................................................ Sorocybe<br />

19. Conidiogenous loci with few, mostly 1–2 conidiogenous loci formed as minute spicules; conidia with narrow hila (shallowly apiculate);<br />

plant pathogenic, causing bud blast <strong>and</strong> twig blight ....................................................................................................................... Seifertia<br />

20(18). Conidiophores unbranched or occasionally branched; conidiogenous cells distinctly inflated, ampulliform, doliiform or clavate, nondenticulate;<br />

conidia at least partly globose, dark brown when mature; colonies effuse, dark; wood-inhabiting .......... Phaeoblastophora<br />

52


<strong>Cladosporium</strong> <strong>and</strong> morphologically <strong>similar</strong> genera<br />

20. Conidiogenous cells not inflated, if somewhat inflated, loci denticle-like or conidia non-globose ............................................................ 21<br />

21. Conidiophores penicillate, i.e., with an unbranched stipe <strong>and</strong> distinct terminal branched “head” composed of branchlets, conidiogenous<br />

cells <strong>and</strong>/or ramoconidia .......................................................................................................................................................................... 22<br />

21. Conidiophores non-penicillate, i.e., irregularly <strong>and</strong> loosely branched, branchings not confined to the apical portion, sometimes only with<br />

short lateral branchlets, or unbranched ................................................................................................................................................... 27<br />

22. Penicillate apex simple, only composed of a single terminal conidiogenous cells giving rise to several ramoconidia which form secondary<br />

ramoconidia <strong>and</strong> conidia ......................................................................................................................... Penidiella p.p. [P. strumelloidea]<br />

22. Penicillate apex more complex, composed of true branchlets <strong>and</strong>/or conidiogenous cells <strong>and</strong> ramoconidia .......................................... 23<br />

23. Conidiophores with a compact, dense, subglobose to broadly ovoid head; conidiogenous loci <strong>and</strong> conidial hila unthickened or almost so,<br />

but distinct by being darkened-refractive [fruiting dimorphic, periconioid branched conidiophores formed on overwintered<br />

stem of Paeonia spp., unbranched cladosporioid conidiophores on leaf spots, biotrophic] (belonging to the Capnodiales)<br />

.................................................................................................................................................................................... Dichocladosporium<br />

23. Penicillate apex looser, neither compact nor subglobose ........................................................................................................................ 24<br />

24. Branched head composed of short branchlets <strong>and</strong> conidiogenous cells; ramoconidia lacking; conidiogenous cells subcylindrical to<br />

subclavate, non-geniculate; conidiogenous loci usually numerous <strong>and</strong> aggregated, terminal <strong>and</strong> lateral, non-protuberant, flat, conspicuous,<br />

thickened <strong>and</strong> darkened, at least around the rim; conidia solitary or in short chains ............................................................... Periconiella<br />

24. Ramoconidia often present; conidiogenous cells distinct, sympodial, somewhat geniculate or subdenticulate; conidiogenous loci<br />

inconspicuous or somewhat protruding, denticle-like, unthickened or almost so, not or somewhat darkened-refractive; conidia in long,<br />

often branched chains ............................................................................................................................................................................. 25<br />

25. Branched apex composed of short branchlets consisting of conidiogenous cells or ramoconidia, in pairs or whorls of 3–4, mostly distinctly<br />

constricted at the base; hyperparasitic on Asterina spp. ................................................................................................. Parapericoniella<br />

25. Branched apex distinct, composed of branchlets, conidiogenous cells <strong>and</strong>/or ramoconidia, if true branchlets lacking conidiogenous cells<br />

<strong>and</strong> ramoconidia not in whorls <strong>and</strong> not distinctly constricted at the base; saprobic or biotrophic ............................................................ 26<br />

26. Penicillate apex of the conidiophores loosely to densely branched, occasionally metula-like, base of the conidiophores simple,<br />

undifferentiated; saprobic or biotrophic (Teratosphaeriaceae, Capnodiales) .............................................................................. Penidiella<br />

26. Penicillate apex always dense, metula-like, base of the conidiophores swollen or lobed, often with rhizoid hyphae; plant pathogenic [on<br />

banana] (Chaetothyriales) ........................................................................................................................................ Metulocladosporiella<br />

27(21). Conidiophores simple or branched; septa of the conidiophores <strong>and</strong> conidia becoming thick-walled <strong>and</strong> dark; conidiogenous loci<br />

subdenticulate, somewhat thickened <strong>and</strong> conspicuously darkened-refractive; cultures producing ample amounts of volatile metabolites<br />

causing skin irritation after exposure to the fungus; saprobic (isolated from mouldy paint) ...................................... Toxicocladosporium<br />

27. Without conspicuously thickened-darkened septa; cultures without irritant, volatile metabolites ............................................................ 28<br />

28. Conidiogenous loci conspicuous, distinctly thickened <strong>and</strong> darkened (visible as small dark circles when viewed upon the scar), sometimes<br />

on small shoulders formed by sympodial proliferation, but not distinctly denticulate (Capnodiales) ....................................................... 29<br />

28. Conidiogenous loci inconspicuous or conspicuous by being denticle-like, not or barely thickened, not darkened or at most upper truncate<br />

end very slightly thickened <strong>and</strong> somewhat darkened-refractive .............................................................................................................. 31<br />

29. Mycelium smooth; conidiophores <strong>and</strong> conidia smooth or almost so, at most faintly rough-walled; conidiophores solitary, fasciculate,<br />

sporodochial to synnematous; biotrophic, usually leaf-spotting (Mycosphaerella anamorphs, Mycosphaerellaceae)<br />

................................................................................................................. Passalora emend. (incl. Mycovellosiella, Phaeoramularia, etc.)<br />

29. At least mycelium distinctly verruculose .................................................................................................................................................. 30<br />

30. Mycelium, conidiophores <strong>and</strong> conidia coarsely verruculose-verrucose; conidial shape variable, often irregular; isolated from a lichen<br />

(Dirina) .................................................................................................................................................................... Verrucocladosporium<br />

30. Mycelium verruculose; conidiophores mostly smooth, sometimes somewhat rough-walled, conidia smooth to distinctly verruculose;<br />

biotrophic, often leaf-spotting ......................................................................................................................................................... Stenella<br />

31(28). Conidiophores with swollen, often lobed base ................................................................................................................................... 32<br />

31. Conidiophores without swollen base, at most slightly swollen, but not lobed .......................................................................................... 35<br />

32. Conidia septate ........................................................................................................................................................................................ 33<br />

32. Conidia aseptate ...................................................................................................................................................................................... 34<br />

33. Conidiophores with a single, terminal, monoblastic, determinate conidiogenous cell giving rise to a single ramoconidium that forms<br />

simple or branched chains of conidia ...................................................................................................................... Parapleurotheciopsis<br />

www.studiesinmycology.org<br />

53


Crous et al.<br />

33. Terminal conidiogenous cells polyblastic, with several denticle-like conidiogenous loci ................. Anungitea p.p. (e.g. A. longicatenata)<br />

34(32). Conidiogenous cells terminal, monoblastic, with a single ramoconidium giving rise to conidial chains or occasionally with 2(–3)<br />

denticle-like loci; base of the conidiophores often with rhizoid hyphae ...................................................................... Rhizocladosporium<br />

34. Conidiogenous cells polyblastic, with two or several denticle-like loci; base of the conidiophores without rhizoid hyphae<br />

...................................................................................................................................................................................... Subramaniomyces<br />

35(31). Conidiophores unbranched, with a terminal monoblastic conidiogenous cell, determinate or percurrent ......................................... 36<br />

35. Conidiophores branched or unbranched, but conidiogenous cells at least partly polyblastic .................................................................. 38<br />

36. Conidiogenous cell giving rise to a single ramoconidium which forms simple or branched chains of 0(–1)-septate conidia<br />

................................................................................................................................................ Parapleurotheciopsis p.p. (P. coccolobae)<br />

36. Conidiogenous cells giving rise to simple conidial chains without ramoconidia; conidia septate ............................................................ 37<br />

37. Conidiophores sometimes with percurrent proliferations; conidiophores <strong>and</strong> conidia with somewhat thickened, dark walls; conidia 1–10-<br />

septate, width usually exceeding 4 µm ................................................................................................................................ Heteroconium<br />

37. Percurrent proliferations lacking; conidiophores <strong>and</strong> conidia delicate, thin-walled <strong>and</strong> paler; conidia usually 0–1(–3)-septate <strong>and</strong> narrow,<br />

usually below 4 µm wide (Chaetothyriales) ......................................................................... Cladophialophora p.p. (e.g. C. chaetospira)<br />

[<strong>similar</strong> anamorphs of the Venturiaceae, see Fusicladium (incl. Pseudocladosporium)]<br />

38(35). Conidiophores often branched; conidiogenous loci distinctly denticle-like or subdenticulate; conidia aseptate; lignicolous, on dead<br />

wood, resin or isolated from hydrocarbone-rich substrates (jet-fuel, cosmetics, etc.) ............................................................................. 39<br />

38. Either with unbranched conidiophores or conidiogenous loci not distinctly denticle-like, or conidia septate, or on other substrates<br />

.................................................................................................................................................................................................................. 42<br />

39. Conidiogenous cells distinctly denticulate; conidia rather broad, approx. 7–13 µm ............................................................. Haplotrichum<br />

39. Conidiogenous cells non-denticulate or at most subdenticulate; conidia narrower, approx. 3–6 µm ...................................................... 40<br />

40. Colonies effuse, dense, but felted, black, brittle <strong>and</strong> appearing carbonaceous when dry; conidiophores solitary, brown; conidiogenous cells<br />

terminal <strong>and</strong> pleurogenous; conidia pale to dark brown, lateral walls conspicuously thicker than the hila; on conifer resin<br />

......................................................................................................................... Sorocybe (mononematous form, Hormodendrum resinae)<br />

40. Colonies effuse, dense, resupinate, hypochnoid, powdery, chocolate-brown <strong>and</strong>/or conidiophores lightly pigmented; conidia subhyaline<br />

to lightly pigmented <strong>and</strong>/or lateral walls not thickener than poles; on dead wood or isolated from hydrocarbone-rich substrates (jet-fuel,<br />

cosmetics, etc.) ........................................................................................................................................................................................ 41<br />

41. Colonies effuse, dense, resupinate, hypochnoid, powdery, chocolate-brown; conidiophores smooth; conidia subhyaline to very pale<br />

yellowish, hila very thin; on dead wood ......................................................................................................................... Parahaplotrichum<br />

41. Colonies neither resupinate nor hypochnoid; conidiophores warty; lateral walls of the conidia not thicker than the hila; isolated from<br />

hydrocarbone-rich substrates (jet-fuel, cosmetics, etc.) ......................................................................................................... Hormoconis<br />

42(38). Conidiophores simple or branched; conidiogenous cells monoblastic or occasionally polyblastic; conidiogenous loci subdenticulate,<br />

neither thickened nor darkened, forming simple or branched chains of regular conidia, uniform in shape, size <strong>and</strong> septation<br />

................................................................................................................................................................................................... Septonema<br />

42. Conidia not uniform in shape, size <strong>and</strong> septation; conidiogenous loci flat-tipped, subdenticulate, unthickened or slightly so, not to<br />

somewhat darkened-refractive ................................................................................................................................................................. 43<br />

43. Conidiophores simple or branched; in culture forming abundant chlamydospores; mostly soil-borne <strong>and</strong> heat-resistant (Teratosphaeriaceae,<br />

Capnodiales) ................................................................................................................................................................................ Devriesia<br />

43. Without chlamydospores in culture; phylogenetically distinct .................................................................................................................. 44<br />

44. Conidiophores dimorphic; conidia mostly aseptate, hila inconspicuous, neither thickened nor darkened (Pleosporales)<br />

.................................................................................................................................................................................... Ochrocladosporium<br />

44. Conidiophores either uniform or conidia at least partly septate or hila more conspicuous by being slightly thickened or at least somewhat<br />

darkened or refractive; phylogenetically distinct ...................................................................................................................................... 45<br />

45. Phialidic synanamorphs often present, but sometimes also lacking; saprobic, rarely plant pathogenic, often human pathogenic<br />

(Herpotrichiellaceae, Chaetothyriales) .......................................................................................................................... Cladophialophora<br />

45. Without phialidic synanamorphs; saprobic or plant pathogenic; phylogenetically distinct ....................................................................... 46<br />

46. Conidiophores usually unbranched (Venturia, Venturiaceae) .......................................... Fusicladium s. lat. (incl. Pseudocladosporium)<br />

[<strong>similar</strong>, barely distinguishable taxa, also clustering in the Venturiaceae, but apart from the Venturia clade are tentatively referred to as<br />

Anungitea until this <strong>genus</strong> will be resolved by sequences of its type species]<br />

54


<strong>Cladosporium</strong> <strong>and</strong> morphologically <strong>similar</strong> genera<br />

46. Conidiophores simple to often irregularly branched; conidia delicate, narrow, 1–3 µm wide, hyaline to pale olivaceous (not belonging to<br />

the Venturiaceae) .................................................................................................................................................................. Polyscytalum<br />

Discussion<br />

Phylogenetic studies conducted on species of <strong>Cladosporium</strong> s. lat.<br />

proved the <strong>genus</strong> to be highly heterogeneous (Braun et al. 2003). It<br />

could be demonstrated that various anamorphs, previously referred<br />

to as <strong>Cladosporium</strong>, e.g. <strong>Cladosporium</strong> fulvum Cooke [≡ Passalora<br />

fulva (Cooke) U. Braun & Crous], have to be excluded since they<br />

clustered in the Mycosphaerella clade (Mycosphaerellaceae).<br />

Previous re-examinations <strong>and</strong> reassessments of human pathogenic<br />

<strong>Cladosporium</strong> species, including morphology, biology/ecology,<br />

physiology <strong>and</strong> molecular data (Masclaux et al. 1995, Untereiner<br />

1997, Gerrits van den Ende & de Hoog 1999, Untereiner & Naveau<br />

1999, Untereiner et al. 1999; de Hoog et al. 2000), could also be<br />

confirmed. In all phylogenetic analyses, it could be shown that the<br />

human pathogenic fungi concerned formed a clade belonging to<br />

the Herpotrichiellaceae (Capronia Sacc./Cladophialophora Borelli).<br />

Venturia anamorphs with catenate conidia, previously often<br />

assigned to <strong>Cladosporium</strong> s. lat., clustered together with other<br />

anamorphs of the Venturiaceae, <strong>and</strong> formed a monophyletic clade<br />

(Braun et al. 2003, Schubert et al. 2003, Beck et al. 2005). Venturia<br />

has now also been shown to accommodate less well-known<br />

anamorph genera such as Pseudocladosporium, which represent<br />

an additional synonym of Fusicladium Bonord. (Crous et al. 2007<br />

– this volume).<br />

Seifert et al. (2004) examined morphological, ecological <strong>and</strong><br />

molecular characters of <strong>Cladosporium</strong> staurophorum (W.B. Kendr.)<br />

M.B. Ellis <strong>and</strong> three allied heat-resistant species <strong>and</strong> placed them<br />

in the new <strong>genus</strong> Devriesia, which formed a monophyletic group<br />

apart from the <strong>Cladosporium</strong> clade. Crous et al. (2006b) erected<br />

the <strong>genus</strong> Cladoriella Crous for a saprobic species (incertae sedis)<br />

characterised by having narrowly ellipsoidal to cylindrical or fusoid,<br />

0–1-septate, medium brown, thick-walled, finely verruculose<br />

conidia arranged in simple or branched chains, with thickened,<br />

darkened, refractive hila, with a minute central pore. <strong>Cladosporium</strong><br />

musae E.W. Mason, the causal agent of banana speckle disease,<br />

has recently been shown to be allied to the Chaetothyriales (Crous<br />

et al. 2006a), <strong>and</strong> was placed in a new <strong>genus</strong>, Metulocladosporiella<br />

with C. musae as type species. Digitopodium U. Braun, Heuchert<br />

& K. Schub. (type species: <strong>Cladosporium</strong> hemileiae Steyaert)<br />

<strong>and</strong> Parapericoniella U. Braun, Heuchert & K. Schub. (type<br />

species: <strong>Cladosporium</strong> asterinae Deighton) represent two new<br />

genera of hyperparasitic hyphomycetes, introduced due to unique<br />

morphological features <strong>and</strong> striking differences to <strong>Cladosporium</strong> s.<br />

str. (Heuchert et al. 2005), but have as yet been excluded from<br />

DNA-based studies due to the absence of cultures. Schubert et al.<br />

(2007a – this volume) introduced a new <strong>genus</strong>, Dichocladosporium<br />

K. Schub., U. Braun & Crous (allied to the Davidiellaceae, Capnodiales)<br />

to accommodate a fungus with dimorphic fruiting that is<br />

pathogenic to Paeonia spp. <strong>The</strong> present study introduced yet several<br />

additional cladosporium-like genera, which could be distinguished<br />

based on their morphology <strong>and</strong> distinct DNA phylogeny, namely<br />

Ochrocladosporium (Pleosporales), Rhizocladosporium<br />

(incertae sedis), Rachicladosporium, Toxicocladosporium <strong>and</strong><br />

Verrucocladosporium (Capnodiales).<br />

Although all these genera are cladosporium-like, <strong>and</strong> many<br />

have in the past been confused with <strong>Cladosporium</strong> s. str., the unique<br />

www.studiesinmycology.org<br />

coronate scar type of <strong>Cladosporium</strong> s. str. allows a critical revision<br />

of cladosporioid hyphomycetes, based on reliable, distinctive<br />

morphological characters. In all cases where cladosporium-like<br />

(<strong>Cladosporium</strong> s. lat.) hyphomycetes clearly clustered apart<br />

from <strong>Cladosporium</strong> s. str. in the phylogenetic analyses, it could<br />

be demonstrated that the fungal groups concerned were also<br />

morphologically unambiguously distinguished, above all with regard<br />

to the structure of the conidiogenous hila. Hence, the excluded<br />

groups of species, belonging in other genera, sometimes even<br />

in new genera, are genetically as well as morphologically clearly<br />

distinct from <strong>Cladosporium</strong> s. str.<br />

Acknowledgements<br />

We thank F. Hill, A. Aptroot, F.M. Dugan, R.F. Castañeda <strong>and</strong> W. Gams for providing<br />

collections <strong>and</strong> cultures of <strong>Cladosporium</strong> <strong>and</strong> cladosporium-like species over the<br />

past few years, without which this study would not have been possible. A research<br />

visit of K. Schubert to <strong>CBS</strong> was supported by a Synthesys grant (No. 2559), <strong>and</strong><br />

the Odo van Vloten Stichting. We thank M. Vermaas for preparing the photographic<br />

plates, <strong>and</strong> A. van Iperen for preparing all the fungal cultures for examination. H.-J.<br />

Schroers is thanked for generating some of the sequence data used in this paper.<br />

A.J.L. Phillips is thanked for providing comments on an earlier draft of the paper.<br />

References<br />

Aptroot A (2006). Mycosphaerella <strong>and</strong> its anamorphs: 2. Conspectus of<br />

Mycosphaerella. <strong>CBS</strong> Biodiversity Series 5: 1–231.<br />

Arzanlou M, Groenewald JZ, Gams W, Braun U, Shin H-D, Crous PW (2007).<br />

Phylogenetic <strong>and</strong> morphotaxonomic revision of Ramichloridium <strong>and</strong> allied<br />

genera. Studies in Mycology 58: 57–93.<br />

Arx JA von (1974). <strong>The</strong> Genera of Fungi Sporulating in Pure Culture. 2 nd edn. J.<br />

Cramer, Vaduz.<br />

Arx JA von (1981). <strong>The</strong> Genera of Fungi Sporulating in Pure Culture. 3 rd edn. J.<br />

Cramer, Vaduz.<br />

Arx JA von (1983). Mycosphaerella <strong>and</strong> its anamorphs. Proceedings of the<br />

Koninklijke Nederl<strong>and</strong>se Akademie van Wetenschappen, Ser. C, 86(1): 15–54.<br />

Beck A, Ritschel A, Schubert K, Braun U, Triebel D (2005). Phylogenetic relationship<br />

of the anamorphic <strong>genus</strong> Fusicladium s. lat. as inferred by ITS data. Mycological<br />

Progress 4: 111–116.<br />

Braun U (1995). A monograph of Cercosporella, Ramularia <strong>and</strong> allied genera<br />

(phytopathogenic hyphomycetes) Vol. 1. IHW-Verlag, Eching.<br />

Braun U (1996). Taxonomic notes on some species of the Cercospora complex (IV).<br />

Sydowia 48: 205–217.<br />

Braun U (1998). A monograph of Cercosporella, Ramularia <strong>and</strong> allied genera<br />

(phytopathogenic hyphomycetes) Vol. 2. IHW-Verlag, Eching.<br />

Braun U, Crous PW, Dugan FM, Groenewald JZ, Hoog GS de (2003). Phylogeny<br />

<strong>and</strong> taxonomy of cladosporium-like hyphomycetes, including Davidiella gen.<br />

nov., the teleomorph of <strong>Cladosporium</strong> s. str. Mycological Progress 2: 3–18.<br />

Braun U, Hill CF (2002). Some new micromycetes from New Zeal<strong>and</strong>. Mycological<br />

Progress 1: 19–30.<br />

Brown KB, Hyde KD, Guest DI (1998). Preliminary studies on endophytic fungal<br />

communities of Musa acuminata species complex in Hong Kong <strong>and</strong> Australia.<br />

Fungal Diversity 1: 27–51.<br />

Castañeda RF, Kendrick B (1990). Conidial fungi from Cuba II. University of Waterloo<br />

Biological Series 33: 1–61.<br />

Crous PW (1998). Mycosphaerella spp. <strong>and</strong> their anamorphs associated with leaf<br />

spot diseases of Eucalyptus. Mycologia Memoir 21: 1–170.<br />

Crous PW, Aptroot A, Kang J-C, Braun U, Wingfield MJ (2000). <strong>The</strong> <strong>genus</strong><br />

Mycosphaerella <strong>and</strong> its anamorphs. Studies in Mycology 45: 107–121.<br />

Crous PW, Braun U (2003). Mycosphaerella <strong>and</strong> its anamorphs: 1. Names published<br />

in Cercospora <strong>and</strong> Passalora. <strong>CBS</strong> Biodiversity Series 1: 1–569.<br />

Crous PW, Groenewald JZ, Mansilla JP, Hunter GC, Wingfield MJ (2004).<br />

Phylogenetic reassessment of Mycosphaerella spp. <strong>and</strong> their anamorphs<br />

occurring on Eucalyptus. Studies in Mycology 50: 195–214.<br />

55


Crous et al.<br />

Crous PW, Kang JC, Braun U (2001). A phylogenetic redefinition of anamorph<br />

genera in Mycosphaerella based on ITS rDNA sequences. Mycologia 93:<br />

1081–1101.<br />

Crous PW, Schroers HJ, Groenewald JZ, Braun U, Schubert K (2006a).<br />

Metulocladosporiella gen. nov. for the causal organism of <strong>Cladosporium</strong><br />

speckle disease of banana. Mycological Research 110: 264–275.<br />

Crous PW, Schubert K, Braun U, Hoog GS de, Hocking AD, Shin H-D, Groenewald<br />

JZ (2007). Opportunistic, human-pathogenic species in the Herpotrichiellaceae<br />

are phenotypically <strong>similar</strong> to saprobic or phytopathogenic species in the<br />

Venturiaceae. Studies in Mycology 58: 185–217.<br />

Crous PW, Slippers B, Wingfield MJ, Rheeder J, Marasas WFO, Phillips AJL, Alves<br />

A, Burgess T, Barber P, Groenewald JZ (2006d). Phylogenetic lineages in the<br />

Botryosphaeriaceae. Studies in Mycology 55: 235–253.<br />

Crous PW, Verkley GJM, Groenewald JZ (2006b). Eucalyptus microfungi known<br />

from culture. 1. Cladoriella <strong>and</strong> Fulvoflamma genera nova, with notes on some<br />

other poorly known taxa. Studies in Mycology 55: 53–63.<br />

Crous PW, Wingfield MJ, Mansilla JP, Alfenas AC, Groenewald JZ (2006c).<br />

Phylogenetic reassessment of Mycosphaerella spp. <strong>and</strong> their anamorphs<br />

occurring on Eucalyptus. II. Studies in Mycology 55: 99–131.<br />

David JC (1997). A contribution to the systematics of <strong>Cladosporium</strong>. Revision of the<br />

fungi previously referred to Heterosporium. Mycological Papers 172: 1–157.<br />

Dugan FM, Schubert K, Braun U (2004). Check-list of <strong>Cladosporium</strong> names.<br />

Schlechtendalia 11: 1–103.<br />

Ellis MB (1971). Dematiaceous hyphomycetes. Commonwealth Mycological<br />

Institute, Kew.<br />

Ellis MB (1976). More <strong>dematiaceous</strong> hyphomycetes. Commonwealth Mycological<br />

Institute, Kew.<br />

El-Morsy EM (2000). Fungi isolated from the endorhizosphere of halophytic plants<br />

from the Red Sea Coast of Egypt. Fungal Diversity 5: 43–54<br />

Gams W, Verkley GJM, Crous PW (2007). <strong>CBS</strong> Course of Mycology, 5 th ed.<br />

Centraalbureau voor Schimmelcultures, Utrecht.<br />

Gerrits van den Ende AHG, Hoog GS de (1999). Variability <strong>and</strong> molecular diagnostics<br />

of the neurotropic species Cladophialophora bantiana. Studies in Mycology 43:<br />

151–162.<br />

Goh T-K, Hyde KD (1998). A synopsis of <strong>and</strong> key to Diplococcium species, based on<br />

the literature, with a description of a new species. Fungal Diversity 1: 65–83.<br />

Hawksworth DL (1979). <strong>The</strong> lichenicolous hyphomycetes. Bulletin of the British<br />

Museum (Natural History), Botany 6: 183–300.<br />

Heuchert B, Braun U, Schubert K (2005). Morphotaxonomic revision of fungicolous<br />

<strong>Cladosporium</strong> species (hyphomycetes). Schlechtendalia 13: 1–78.<br />

Ho MH-M, Castañeda RF, Dugan FM, Jong, SC (1999). <strong>Cladosporium</strong> <strong>and</strong><br />

Cladophialophora in culture: descriptions <strong>and</strong> an exp<strong>and</strong>ed key. Mycotaxon<br />

72: 115–157.<br />

Hoog GS de, Batenburg van der Vegte WH (1989). Retroconis, a new <strong>genus</strong> of<br />

ascomycetous hyphomycetes. Studies in Mycology 31: 99–105.<br />

Hoog GS de, Gerrits van den Ende AHG (1998). Molecular diagnostics of clinical<br />

strains of filamentous Basidiomycetes. Mycoses 41: 183–189.<br />

Hoog GS de, Guarro, J, Gené J, Figueras MJ (2000). Atlas of clinical fungi, 2 nd ed.<br />

Centraalbureau voor Schimmelcultures, Utrecht <strong>and</strong> Universitat rovira I virgili,<br />

Reus.<br />

Inacio J, Pereira P, de Cavalho M, Fonseca A, Amaral-Collaco MT, Spencer-Martins<br />

I (2002). Estimation <strong>and</strong> diversity of phylloplane mycobiota on selected plants in<br />

a Mediterranean-type ecosystem in Portugal. Microbial Ecology 44: 344–353.<br />

Islam M, Hasin F (2000). Studies on phylloplane mycoflora of Amaranthus viridis L.<br />

National Academy Science Letters, India 23: 121–123.<br />

Jager ES de, Wehner FC, Karsten L (2001). Microbial ecology of the mango<br />

phylloplane. Microbial Ecology 42: 201–207.<br />

Kirk PM (1982). New or interesting microfungi. IV. Dematiaceous hyphomycetes<br />

from Devon. Transactions of the Britisch Mycological Society 78: 55–74.<br />

Levetin E, Dorsey K (2006). Contribution to leaf surface fungi to the air spora.<br />

Aerobiologia 22: 3–12.<br />

Masclaux F, Guého E, Hoog GS de, Christen R (1995). Phylogenetic relationship of<br />

human-pathogenic <strong>Cladosporium</strong> (Xylohypha) species. Journal of Medical <strong>and</strong><br />

Veterinary Mycology 33: 327–338.<br />

Matsushima T (1975). Icones microfungorum a Matsushima lectorum. Kobe.<br />

McGinnis MR, Padhye AA (1978). <strong>Cladosporium</strong> castellanii is a synonym of Stenella<br />

araguata. Mycotaxon 7: 415–418.<br />

McKemy JM, Morgan-Jones G (1990). Studies in the <strong>genus</strong> <strong>Cladosporium</strong> sensu<br />

lato II. Concerning Heterosporium gracile, the causal organism of leaf spot<br />

disease of Iris species. Mycotaxon 39: 425–440.<br />

Minoura K (1966). Taxonomic studies on cladosporia (IV). Morphological properties<br />

(part II). Journal of Fermentation Technology 44: 137–149.<br />

Moncalvo J-M, Rehner SA, Vilgalys R (1993). Systematics of Lyophyllum section<br />

Difformia based on evidence from culture studies <strong>and</strong> ribosomal DNA<br />

sequences. Mycologia 85: 788–794.<br />

Park HG, Managbanag JR, Stamenova EK, Jong SC (2004). Comparative analysis<br />

of common indoor <strong>Cladosporium</strong> species based on molecular data <strong>and</strong> conidial<br />

characters. Mycotaxon 89: 441–451.<br />

Partridge EC, Morgan-Jones G (2002). Notes on hyphomycetes, LXXXVIII. New<br />

genera in which to classify Alysidium resinae <strong>and</strong> Pycnostysanus azaleae, with<br />

a consideration of Sorocybe. Mycotaxon 83: 335–352.<br />

Rayner RW (1970). A mycological colour chart. CMI <strong>and</strong> British Mycological Society.<br />

Kew.<br />

Rehner SA, Samuels GJ (1994). Taxonomy <strong>and</strong> phylogeny of Gliocladium analysed<br />

from nuclear large subunit ribosomal DNA sequences. Mycological Research<br />

98: 625–634.<br />

Riesen T, Sieber T (1985). Endophytic fungi in winter wheat (Triticum aestivum L.).<br />

Swiss Federal Institute of Technology, Zürich.<br />

Roquebert MF (1981). Analyse des phénoménes pariétaux au cours de la<br />

conidiogenése chez quelques champignon microscopiques. Memoires du<br />

Museum d’Histoire Naturelle, Sér B, Botanique 28: 3–79.<br />

Schoch C, Shoemaker RA, Seifert KA, Hambleton S, Spatafora JW, Crous PW<br />

(2006). A multigene phylogeny of the Dothideomycetes using four nuclear loci.<br />

Mycologia 98: 1041–1052.<br />

Schubert K (2005a). Morphotaxonomic revision of foliicolous <strong>Cladosporium</strong> species<br />

(hyphomycetes). PhD thesis, Martin-Luther-University, Halle.<br />

Schubert K (2005b). Taxonomic revision of the <strong>genus</strong> <strong>Cladosporium</strong> s. lat. 3. A<br />

revision of <strong>Cladosporium</strong> species described by J.J. Davis <strong>and</strong> H.C. Greene<br />

(WIS). Mycotaxon 92: 55–76.<br />

Schubert K, Braun U (2004). Taxonomic revision of the <strong>genus</strong> <strong>Cladosporium</strong> s. lat. 2.<br />

Morphotaxonomic examination of <strong>Cladosporium</strong> species occurring on hosts of<br />

the families Bignoniaceae <strong>and</strong> Orchidaceae. Sydowia 56: 76–97.<br />

Schubert K, Braun U (2005a). Taxonomic revision of the <strong>genus</strong> <strong>Cladosporium</strong><br />

s. lat. 1. Species reallocated to Fusicladium, Parastenella, Passalora,<br />

Pseudocercospora <strong>and</strong> Stenella. Mycological Progress 4: 101–109.<br />

Schubert K, Braun U (2005b). Taxonomic revision of the <strong>genus</strong> <strong>Cladosporium</strong> s.<br />

lat. 4. Species reallocated to Asperisporium, Dischloridium, Fusicladium,<br />

Passalora, Pseudasperisporium <strong>and</strong> Stenella. Fungal Diversity 20: 187–208.<br />

Schubert K, Braun U (2007). Taxonomic revision of the <strong>genus</strong> <strong>Cladosporium</strong> s. lat.<br />

6. New species, reallocations to <strong>and</strong> synonyms of Cercospora, Fusicladium,<br />

Passalora, Septonema <strong>and</strong> Stenella. Nova Hedwigia 84: 189–208.<br />

Schubert K, Braun U, Groenewald JZ, Crous PW (2007a). <strong>Cladosporium</strong> leaf-blotch<br />

<strong>and</strong> stem rot of Paeonia spp. caused by Dichocladosporium chlorocephalum<br />

gen. nov. Studies in Mycology 58: 95–104.<br />

Schubert K, Braun U, Mułenko W (2006). Taxonomic revision of the <strong>genus</strong><br />

<strong>Cladosporium</strong> s. lat. 5. Validations <strong>and</strong> descriptions of new species.<br />

Schlechtendalia 14: 55–83.<br />

Schubert K, Groenewald JZ, Braun U, Dijksterhuis J, Starink M, Hill CF, Zalar P,<br />

Hoog GS de, Crous PW (2007b). Biodiversity in the <strong>Cladosporium</strong> herbarum<br />

complex (Davidiellaceae, Capnodiales), with st<strong>and</strong>ardisation of methods for<br />

<strong>Cladosporium</strong> taxonomy <strong>and</strong> diagnostics. Studies in Mycology 58: 105–156.<br />

Schubert K, Ritschel A, Braun U (2003). A monograph of Fusicladium s. lat.<br />

(hyphomycetes). Schlechtendalia 9: 1–132.<br />

Seifert K, Nickerson NL, Corlett M, Jackson ED, Lois-Seize G, Davies RJ (2004).<br />

Devriesia, a new hyphomycete <strong>genus</strong> to accommodate heat-resistant,<br />

cladosporium-like fungi. Canadian Journal of Botany 82: 914–926.<br />

Stohr SN, Dighton J (2004). Effects of species diversity on establishment <strong>and</strong><br />

coexistence: A phylloplane fungal community model system. Microbial Ecology<br />

48: 431–438.<br />

Untereiner WA (1997). Taxonomy of selected members of the ascomycete <strong>genus</strong><br />

Capronia with notes on anamorph-teleomorph connections. Mycologia 89:<br />

120–131.<br />

Untereiner WA, Gerrits van den Ende AHG, Hoog GS de (1999). Nutritional<br />

physiology of species of Capronia. Studies in Mycology 43: 98–106.<br />

Untereiner WA, Naveau FA (1999). Molecular systematics of the Herpotrichiellaceae<br />

with an assessment of the phylogenetic positions of Exophiala dermatitidis <strong>and</strong><br />

Phialophora americana. Mycologia 91: 67–83.<br />

Varghese KIM, Rao VG (1979). Forest microfungi – I. Subramaniomyces, a new<br />

<strong>genus</strong> of Hyphomycetes. Kavaka 7: 83–85.<br />

Vilgalys R, Hester M (1990). Rapid genetic identification <strong>and</strong> mapping of<br />

enzymatically amplified ribosomal DNA from several Cryptococcus species.<br />

Journal of Bacteriology 172: 4238–4246.<br />

Vries GA de (1952). Contribution to the Knowledge of the Genus <strong>Cladosporium</strong> Link<br />

ex Fr. Centraalbureau voor Schimmelcultures, Baarn.<br />

White TJ, Bruns T, Lee S, Taylor J (1990). Amplification <strong>and</strong> direct sequencing of<br />

fungal ribosomal RNA genes for phylogenetics. In: PCR Protocols: a guide to<br />

methods <strong>and</strong> applications (Innis MA, Gelf<strong>and</strong> DH, Sninsky JJ, White TJ, eds).<br />

Academic Press, San Diego, California: 315–322.<br />

Wirsel SGR, Runge-Froböse C, Ahrén DG, Kemen E, Oliver RP, Mendgen<br />

KW (2002). Four or more species of <strong>Cladosporium</strong> sympatrically colonize<br />

Phragmites australis. Fungal Genetics <strong>and</strong> Biology 35: 99–113.<br />

56


available online at www.studiesinmycology.org<br />

doi:10.3114/sim.2007.58.03<br />

Studies in Mycology 58: 57–93. 2007.<br />

Phylogenetic <strong>and</strong> morphotaxonomic revision of Ramichloridium <strong>and</strong> allied<br />

genera<br />

M. Arzanlou 1,2 , J.Z. Groenewald 1 , W. Gams 1 , U. Braun 3 , H.-D Shin 4 <strong>and</strong> P.W. Crous 1,2*<br />

1<br />

<strong>CBS</strong> Fungal Biodiversity Centre, Uppsalalaan 8, 3584 CT Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; 2 Laboratory of Phytopathology, Wageningen University, Binnenhaven 5, 6709 PD<br />

Wageningen, <strong>The</strong> Netherl<strong>and</strong>s; 3 Martin-Luther-Universität, Institut für Biologie, Geobotanik und Botanischer Garten, Neuwerk 21, D-06099 Halle (Saale), Germany; 4 Division of<br />

Environmental Science & Ecological Engineering, Korea University, Seoul 136-701, Korea<br />

*Correspondence: Pedro W. Crous, p.crous@cbs.knaw.nl<br />

Abstract: <strong>The</strong> phylogeny of the genera Periconiella, Ramichloridium, Rhinocladiella <strong>and</strong> Veronaea was explored by means of partial sequences of the 28S (LSU) rRNA<br />

gene <strong>and</strong> the ITS region (ITS1, 5.8S rDNA <strong>and</strong> ITS2). Based on the LSU sequence data, ramichloridium-like species segregate into eight distinct clusters. <strong>The</strong>se include the<br />

Capnodiales (Mycosphaerellaceae <strong>and</strong> Teratosphaeriaceae), the Chaetothyriales (Herpotrichiellaceae), the Pleosporales, <strong>and</strong> five ascomycete clades with uncertain affinities.<br />

<strong>The</strong> type species of Ramichloridium, R. apiculatum, together with R. musae, R. biverticillatum, R. cerophilum, R. verrucosum, R. pini, <strong>and</strong> three new species isolated from<br />

Strelitzia, Musa <strong>and</strong> forest soil, respectively, reside in the Capnodiales clade. <strong>The</strong> human-pathogenic species R. mackenziei <strong>and</strong> R. basitonum, together with R. fasciculatum<br />

<strong>and</strong> R. anceps, cluster with Rhinocladiella (type species: Rh. atrovirens, Herpotrichiellaceae, Chaetothyriales), <strong>and</strong> are allocated to this <strong>genus</strong>. Veronaea botryosa, the type<br />

species of the <strong>genus</strong> Veronaea, also resides in the Chaetothyriales clade, whereas Veronaea simplex clusters as a sister taxon to the Venturiaceae (Pleosporales), <strong>and</strong> is placed<br />

in a new <strong>genus</strong>, Veronaeopsis. Ramichloridium obovoideum clusters with Carpoligna pleurothecii (anamorph: Pleurothecium sp., Chaetosphaeriales), <strong>and</strong> a new combination<br />

is proposed in Pleurothecium. Other ramichloridium-like clades include R. subulatum <strong>and</strong> R. epichloës (incertae sedis, Sordariomycetes), for which a new <strong>genus</strong>, Radulidium<br />

is erected. Ramichloridium schulzeri <strong>and</strong> its varieties are placed in a new <strong>genus</strong>, Myrmecridium (incertae sedis, Sordariomycetes). <strong>The</strong> <strong>genus</strong> Pseudovirgaria (incertae sedis)<br />

is introduced to accommodate ramichloridium-like isolates occurring on various species of rust fungi. A veronaea-like isolate from Bertia moriformis with phylogenetic affinity to<br />

the Annulatascaceae (Sordariomycetidae) is placed in a new <strong>genus</strong>, Rhodoveronaea. Besides Ramichloridium, Periconiella is also polyphyletic. Thysanorea is introduced to<br />

accommodate Periconiella papuana (Herpotrichiellaceae), which is unrelated to the type species, P. velutina (Mycosphaerellaceae).<br />

Taxonomic novelties: Myrmecridium Arzanlou, W. Gams & Crous, gen. nov., Myrmecridium flexuosum (de Hoog) Arzanlou, W. Gams & Crous, comb. et stat. nov., Myrmecridium<br />

schulzeri (Sacc.) Arzanlou, W. Gams & Crous var. schulzeri, comb. nov., Myrmecridium schulzeri var. tritici (M.B. Ellis) Arzanlou, W. Gams & Crous, comb. nov., Periconiella<br />

arcuata Arzanlou, S. Lee & Crous, sp. nov., Periconiella levispora Arzanlou, W. Gams & Crous, sp. nov., Pleurothecium obovoideum (Matsush.) Arzanlou & Crous, comb. nov.,<br />

Pseudovirgaria H.D. Shin, U. Braun, Arzanlou & Crous, gen. nov., Pseudovirgaria hyperparasitica H.D. Shin, U. Braun, Arzanlou & Crous, sp. nov., Radulidium Arzanlou, W.<br />

Gams & Crous, gen. nov., Radulidium epichloës (Ellis & Dearn.) Arzanlou, W. Gams & Crous, comb. nov., Radulidium subulatum (de Hoog) Arzanlou, W. Gams & Crous, comb.<br />

nov., Ramichloridium australiense Arzanlou & Crous, sp. nov., Ramichloridium biverticillatum Arzanlou & Crous, nom. nov., Ramichloridium brasilianum Arzanlou & Crous, sp.<br />

nov., Ramichloridium strelitziae Arzanlou, W. Gams & Crous, sp. nov., Rhinocladiella basitona (de Hoog) Arzanlou & Crous, comb. nov., Rhinocladiella fasciculata (V. Rao &<br />

de Hoog) Arzanlou & Crous, comb. nov., Rhinocladiella mackenziei (C.K. Campb. & Al-Hedaithy) Arzanlou & Crous, comb. nov., Rhodoveronaea Arzanlou, W. Gams & Crous,<br />

gen. nov., Rhodoveronaea varioseptata Arzanlou, W. Gams & Crous, sp. nov., Thysanorea Arzanlou, W. Gams & Crous, gen. nov., Thysanorea papuana (Aptroot) Arzanlou, W.<br />

Gams & Crous, comb. nov., Veronaea japonica Arzanlou, W. Gams & Crous, sp. nov., Veronaeopsis Arzanlou & Crous, gen. nov., Veronaeopsis simplex (Papendorf) Arzanlou<br />

& Crous, comb.nov.<br />

Key words: Capnodiales, Chaetothyriales, Mycosphaerella, Periconiella, phylogeny, Rhinocladiella, Veronaea.<br />

Introduction<br />

<strong>The</strong> anamorph <strong>genus</strong> Ramichloridium Stahel ex de Hoog 1977<br />

presently accommodates a wide range of species with erect, dark,<br />

more or less differentiated, branched or unbranched conidiophores<br />

<strong>and</strong> predominantly aseptate conidia produced on a sympodially<br />

proliferating rachis (de Hoog 1977). This heterogeneous group<br />

of anamorphic fungi includes species with diverse life styles, viz.<br />

saprobes, human <strong>and</strong> plant pathogens, most of which were classified<br />

by Schol-Schwarz (1968) in Rhinocladiella Nannf. according to a<br />

very broad generic concept. Ramichloridium was originally erected<br />

by Stahel (1937) with R. musae Stahel as type species. However,<br />

because his publication lacked a Latin diagnosis, the <strong>genus</strong> was<br />

invalid. Stahel also invalidly described Chloridium musae Stahel for<br />

a fungus causing leaf spots (tropical speckle disease) on banana.<br />

Ellis (1976) validated Chloridium musae as Veronaea musae M.B.<br />

Ellis, <strong>and</strong> Ramichloridium musae as Periconiella musae Stahel ex<br />

M.B. Ellis.<br />

Periconiella Sacc. (1885) [type species P. velutina (G. Winter)<br />

Sacc.] differs from Veronaea Cif. & Montemart. chiefly based<br />

on its dark brown, apically branched conidiophores. However,<br />

de Hoog (1977) observed numerous specimens of V. musae to<br />

exhibit branched conidiophores in culture, as did Stahel (1937)<br />

for Ramichloridium musae. De Hoog (1977) subsequently reintroduced<br />

Ramichloridium, but typified it with R. apiculatum (J.H.<br />

Mill., Giddens & A.A. Foster) de Hoog. He regarded V. musae <strong>and</strong><br />

P. musae to be conspecific, <strong>and</strong> applied the name R. musae (Stahel<br />

ex M.B. Ellis) de Hoog to both species, regarding Periconiella<br />

musae as basionym. <strong>The</strong> circumscription by de Hoog was based<br />

on their <strong>similar</strong> morphology <strong>and</strong> ecology. Central in his <strong>genus</strong><br />

concept was the observed presence of more or less differentiated<br />

<strong>and</strong> pigmented conidiophores, with predominantly aseptate conidia<br />

produced on a sympodially proliferating rachis. De Hoog (1977)<br />

also used some ecological features as additional characters to<br />

discriminate Ramichloridium from other genera, noting, for instance,<br />

that species in Ramichloridium were non-pathogenic to humans<br />

(de Hoog 1977, Campbell & Al-Hedaithy 1993). This delimitation,<br />

however, was not commonly accepted (McGinnis & Schell 1980).<br />

De Hoog et al. (1983) further discussed the problematic separation<br />

of Ramichloridium from genera such as Rhinocladiella, Veronaea<br />

<strong>and</strong> <strong>Cladosporium</strong> Link. It was further noted that the main feature to<br />

distinguish Ramichloridium from Rhinocladiella, was the presence<br />

of exophiala-type budding cells in species of Rhinocladiella (de<br />

Hoog 1977, de Hoog et al. 1983, Veerkamp & Gams 1983). <strong>The</strong><br />

separation of Veronaea from this complex is more problematic, as<br />

the circumscriptions provided by Ellis (1976) <strong>and</strong> Morgan-Jones<br />

(1979, 1982) overlap with that of Ramichloridium sensu de Hoog<br />

(1977). <strong>Cladosporium</strong> is more distinct, having very conspicuous,<br />

protuberant, darkened <strong>and</strong> thickened, coronate conidial scars, <strong>and</strong><br />

catenate conidia (David 1997, Braun et al. 2003, Schubert et al.<br />

2007 – this volume).<br />

57


Arzanlou et al.<br />

To date 26 species have been named in Ramichloridium; they<br />

not only differ in morphology, but also in life style. Ramichloridium<br />

mackenziei C.K. Campb. & Al-Hedaithy is a serious human<br />

pathogen, causing cerebral phaeohyphomycosis (Al-Hedaithy et al.<br />

1988, Campbell & Al-Hedaithy 1993), whereas R. musae causes<br />

tropical speckle disease on members of the Musaceae (Stahel 1937,<br />

Jones 2000). Another plant-pathogenic species, R. pini de Hoog &<br />

Rahman, causes a needle disease on Pinus contorta (de Hoog et<br />

al. 1983). Other clinically relevant species of Ramichloridium are R.<br />

basitonum de Hoog <strong>and</strong> occasionally R. schulzeri (Sacc.) de Hoog,<br />

while the remaining species tend to be common soil saprobes.<br />

No teleomorph has thus far been linked to species of<br />

Ramichloridium. <strong>The</strong> main question that remains is whether shared<br />

morphology among the species in this <strong>genus</strong> reflects common<br />

ancestry (Seifert 1993, Untereiner & Naveau 1999). To delineate<br />

anamorphic genera adequately, morphology <strong>and</strong> conidial ontogeny<br />

alone are no longer satisfactory (Crous et al. 2006a, b), <strong>and</strong> DNA<br />

data provide additional characters to help delineate species <strong>and</strong><br />

genera (Taylor et al. 2000, Mostert et al. 2006, Zipfel et al. 2006).<br />

<strong>The</strong> aim of the present study was to integrate morphological<br />

<strong>and</strong> cultural features with DNA sequence data to resolve the<br />

species concepts <strong>and</strong> generic limits of the taxa currently placed in<br />

Periconiella, Ramichloridium, Rhinocladiella <strong>and</strong> Veronaea, <strong>and</strong> to<br />

resolve the status of several new cultures that were isolated during<br />

the course of this study.<br />

Materials <strong>and</strong> Methods<br />

Isolates<br />

Species names, substrates, geographical origins <strong>and</strong> GenBank<br />

accession numbers of the isolates included in this study are listed<br />

in Table 1. Fungal isolates are maintained in the culture collection<br />

of the Centraalbureau voor Schimmelcultures (<strong>CBS</strong>) in Utrecht, the<br />

Netherl<strong>and</strong>s.<br />

DNA extraction, amplification <strong>and</strong> sequence analysis<br />

Genomic DNA was extracted from colonies grown on 2 % malt<br />

extract agar (MEA, Difco) (Gams et al. 2007) using the FastDNA kit<br />

(BIO101, Carlsbad, CA, U.S.A.). <strong>The</strong> primers ITS1 <strong>and</strong> ITS4 (White<br />

et al. 1990) were used to amplify the internal transcribed spacer<br />

region (ITS) of the nuclear ribosomal RNA operon, including: the<br />

3’ end of the 18S rRNA gene, the first internal transcribed spacer<br />

region (ITS1), the 5.8S rRNA gene, the second internal transcribed<br />

spacer region (ITS2) <strong>and</strong> the 5’ end of 28S rRNA gene. Part of<br />

the large subunit 28S rRNA (LSU) gene was amplified with primers<br />

LR0R (Rehner & Samuels 1994) <strong>and</strong> LR5 (Vilgalys & Hester 1990).<br />

<strong>The</strong> ITS region was sequenced only for those isolates for which<br />

these data were not available. <strong>The</strong> ITS analyses confirmed the<br />

proposed classification based on LSU analysis for each major<br />

clade <strong>and</strong> are not presented here in detail; but the sequences are<br />

deposited in GenBank where applicable. <strong>The</strong> PCR reaction was<br />

performed in a mixture with 0.5 units Taq polymerase (Bioline,<br />

London, U.K.), 1× PCR buffer, 0.5 mM MgCl 2<br />

, 0.2 mM of each<br />

dNTP, 5 pmol of each primer, approximately 10–15 ng of fungal<br />

genomic DNA, with the total volume adjusted to 25 µL with<br />

sterile water. Reactions were performed on a GeneAmp PCR<br />

System 9700 (Applied Biosystems, Foster City, CA) with cycling<br />

conditions consisting of 5 min at 96 °C for primary denaturation,<br />

followed by 36 cycles at 96 °C (30 s), 52 °C (30 s), <strong>and</strong> 72 °C<br />

(60 s), with a final 7 min extension step at 72 °C to complete the<br />

reaction. <strong>The</strong> amplicons were sequenced using BigDye Terminator<br />

v. 3.1 (Applied Biosystems, Foster City, CA) or DYEnamicET<br />

Terminator (Amersham Biosciences, Freiburg, Germany) Cycle<br />

Sequencing Kits <strong>and</strong> analysed on an ABI Prism 3700 (Applied<br />

Biosystems, Foster City, CA) under conditions recommended by<br />

the manufacturer. Newly generated sequences were subjected<br />

to a Blast search of the NCBI databases, sequences with high<br />

<strong>similar</strong>ity were downloaded from GenBank <strong>and</strong> comparisons<br />

were made based on the alignment of the obtained sequences.<br />

Sequences from GenBank were also selected for <strong>similar</strong> taxa. <strong>The</strong><br />

LSU tree was rooted using sequences of Athelia epiphylla Pers.<br />

<strong>and</strong> Paullicorticium ansatum Liberta as outgroups. Phylogenetic<br />

analysis was performed with PAUP (Phylogenetic Analysis Using<br />

Parsimony) v. 4.0b10 (Swofford 2003), using the neighbour-joining<br />

algorithm with the uncorrected (“p”), the Kimura 2-parameter <strong>and</strong><br />

the HKY85 substitution models. Alignment gaps longer than 10<br />

bases were coded as single events for the phylogenetic analyses;<br />

the remaining gaps were treated as missing data. Any ties were<br />

broken r<strong>and</strong>omly when encountered. Phylogenetic relationships<br />

were also inferred with the parsimony algorithm using the heuristic<br />

search option with simple taxon additions <strong>and</strong> tree bisection <strong>and</strong><br />

reconstruction (TBR) as the branch-swapping algorithm; alignment<br />

gaps were treated as a fifth character state <strong>and</strong> all characters were<br />

unordered <strong>and</strong> of equal weight. Branches of zero length were<br />

collapsed <strong>and</strong> all multiple, equally parsimonious trees were saved.<br />

Only the first 5 000 equally most parsimonious trees were saved.<br />

Other measures calculated included tree length, consistency<br />

index, retention index <strong>and</strong> rescaled consistency index (TL, CI, RI<br />

<strong>and</strong> RC, respectively). <strong>The</strong> robustness of the obtained trees was<br />

evaluated by 1 000 bootstrap replications. Bayesian analysis was<br />

performed following the methods of Crous et al. (2006c). <strong>The</strong> best<br />

nucleotide substitution model was determined using MrModeltest<br />

v. 2.2 (Nyl<strong>and</strong>er 2004). MrBayes v. 3.1.2 (Ronquist & Huelsenbeck<br />

2003) was used to perform phylogenetic analyses, using a general<br />

time-reversible (GTR) substitution model with inverse gamma<br />

rates, dirichlet base frequencies <strong>and</strong> the temp value set to 0.5. New<br />

sequences were lodged with NCBI’s GenBank (Table 1) <strong>and</strong> the<br />

alignment <strong>and</strong> trees with TreeBASE (www.treebase.org).<br />

Morphology<br />

Cultural growth rates <strong>and</strong> morphology were recorded from colonies<br />

grown on MEA for 2 wk at 24 ºC in the dark, <strong>and</strong> colony colours<br />

were determined by reference to the colour charts of Rayner (1970).<br />

Microscopic observations were made from colonies cultivated on<br />

MEA <strong>and</strong> OA (oatmeal agar, Gams et al. 2007), using a slide culture<br />

technique. Slide cultures were set up in Petri dishes containing 2<br />

mL of sterile water, into which a U-shaped glass rod was placed,<br />

extending above the water surface. A block of freshly growing<br />

fungal colony, approx. 1 cm square was placed onto a sterile<br />

microscope slide, covered with a somewhat larger, sterile glass<br />

cover slip, <strong>and</strong> incubated in the moist chamber. Fungal sporulation<br />

was monitored over time, <strong>and</strong> when optimal, images were captured<br />

by means of a Nikon camera system (Digital Sight DS-5M, Nikon<br />

Corporation, Japan). Structures were mounted in lactic acid, <strong>and</strong><br />

30 measurements (× 1 000 magnification) determined wherever<br />

possible, with the extremes of spore measurements given in<br />

parentheses.<br />

58


Ramichloridium <strong>and</strong> allied genera<br />

Table 1. Isolates of Ramichloridium <strong>and</strong> <strong>similar</strong> genera used for DNA analysis <strong>and</strong> morphological studies.<br />

Species Accession number 1 Source Origin GenBank numbers<br />

(LSU, ITS)<br />

Myrmecridium flexuosum <strong>CBS</strong> 398.76*; IMI 203547 Soil Suriname EU041825, EU041768<br />

Myrmecridium schulzeri <strong>CBS</strong> 100.54; JCM 6974 Soil Zaire EU041826, EU041769<br />

<strong>CBS</strong> 134.68; ATCC 16310 Soil Germany EU041827, EU041770<br />

<strong>CBS</strong> 156.63 Homo sapiens Netherl<strong>and</strong>s EU041828, EU041771<br />

<strong>CBS</strong> 188.96 Soil Papua New Guinea EU041829, EU041772<br />

<strong>CBS</strong> 304.73 Wheat straw South Africa EU041830, EU041773<br />

<strong>CBS</strong> 305.73; JCM 6967 Wheat straw South Africa EU041831, EU041774<br />

<strong>CBS</strong> 325.74; JCM 7234 Triticum aestivum Netherl<strong>and</strong>s EU041832, EU041775<br />

<strong>CBS</strong> 381.87 — Australia EU041833, EU041776<br />

<strong>CBS</strong> 642.76 Malus sylvestris Switzerl<strong>and</strong> EU041834, EU041777<br />

<strong>CBS</strong> 114996 Cannomois virgata South Africa EU041835, EU041778<br />

Periconiella arcuata <strong>CBS</strong> 113477* Ischyrolepsis subverticellata South Africa EU041836, EU041779<br />

Periconiella levispora <strong>CBS</strong> 873.73* Turpinia pomifera Sri Lanka EU041837, EU041780<br />

Periconiella velutina <strong>CBS</strong> 101948*; CPC 2262 Brabejum stellatifolium South Africa EU041838, EU041781<br />

<strong>CBS</strong> 101949; CPC 2263 Brabejum stellatifolium South Africa EU041839, EU041782<br />

<strong>CBS</strong> 101950; CPC 2264 Brabejum stellatifolium South Africa EU041840, EU041783<br />

Pleurothecium obovoideum <strong>CBS</strong> 209.95*; MFC 12477 Pasania edulis Japan EU041841, EU041784<br />

Pseudovirgaria hyperparasitica <strong>CBS</strong> 121735; CPC 10702 On Phragmidium sp. on Rubus coreanus Korea EU041822, EU041765<br />

<strong>CBS</strong> 121738; CPC 10704 On Phragmidium sp. on Rubus coreanus Korea EU041823, EU041766<br />

<strong>CBS</strong> 121739*; CPC 10753<br />

On Pucciniastrum agrimoniae on<br />

Korea<br />

EU041824, EU041767<br />

Agrimonia pilosa<br />

Radulidium epichloës <strong>CBS</strong> 361.63*; MUCL 3124 Epichloë typhina U.S.A. EU041842, EU041785<br />

Radulidium sp. <strong>CBS</strong> 115704 Poaceae Guyana EU041843, EU041786<br />

Radulidium subulatum <strong>CBS</strong> 287.84 Puccinia allii U.K. EU041844, EU041787<br />

Ramichloridium apiculatum<br />

<strong>CBS</strong> 405.76* Phragmites australis Czech Republic EU041845, EU041788<br />

<strong>CBS</strong> 912.96 Incubator for cell cultures Germany EU041846, EU041789<br />

<strong>CBS</strong> 101010 Lasioptera arundinis Czech Republic EU041847, EU041790<br />

<strong>CBS</strong> 156.59*; ATCC 13211; IMI Forest soil U.S.A. EU041848, EU041791<br />

100716; JCM 6972; MUCL 7991;<br />

MUCL 15753; QM 7716<br />

<strong>CBS</strong> 390.67 Cucumis sativus South Africa EU041849, EU041792<br />

<strong>CBS</strong> 391.67; JCM 6966 Aloe sp. South Africa EU041850, EU041793<br />

<strong>CBS</strong> 400.76; IMI 088021 Soil Pakistan EU041851, EU041794<br />

Ramichloridium australiense <strong>CBS</strong> 121710 Musa banksii Australia EU041852, EU041795<br />

Ramichloridium biverticillatum <strong>CBS</strong> 335.36 Musa sapientum — EU041853, EU041796<br />

Ramichloridium brasilianum <strong>CBS</strong> 283.92* Forest soil Brazil EU041854, EU041797<br />

Ramichloridium cerophilum <strong>CBS</strong> 103.59* Sasa sp. Japan EU041855, EU041798<br />

Ramichloridium indicum <strong>CBS</strong> 171.96 — — EU041856, EU041799<br />

Ramichloridium musae <strong>CBS</strong> 190.63; MUCL 9557 Musa sapientum — EU041857, EU041800<br />

<strong>CBS</strong> 365.36*; JCM 6973; MUCL 9556 Musa sapientum Surinam EU041858, EU041801<br />

Ramichloridium pini <strong>CBS</strong> 461.82*; MUCL 28942 Pinus contorta U.K. EU041859, EU041802<br />

Ramichloridium strelitziae <strong>CBS</strong> 121711 Strelitzia sp. South Africa EU041860, EU041803<br />

Rhinocladiella anceps<br />

<strong>CBS</strong> 157.54; ATCC 15680; MUCL Fagus sylvatica France EU041861, EU041804<br />

1081; MUCL 7992; MUCL 15756<br />

<strong>CBS</strong> 181.65*; ATCC 18655; DAOM Soil Canada EU041862, EU041805<br />

84422; IMI 134453; MUCL 8233; OAC<br />

10215<br />

Rhinocladiella basitona <strong>CBS</strong> 101460*; IFM 47593 Homo sapiens Japan EU041863, EU041806<br />

Rhinocladiella fasciculata <strong>CBS</strong> 132.86* Decayed wood India EU041864, EU041807<br />

Rhinocladiella mackenziei <strong>CBS</strong> 367.92; NCPF 2738; UTMB 3169 Homo sapiens Israel EU041865, EU041808<br />

www.studiesinmycology.org<br />

59


Arzanlou et al.<br />

Table 1. (Continued).<br />

Species Accession number 1 Source Origin GenBank numbers<br />

(LSU, ITS)<br />

<strong>CBS</strong> 368.92; UTMB 3170 Homo sapiens Israel EU041866, EU041809<br />

<strong>CBS</strong> 102590; NCPF 2853 Homo sapiens United Arab Emirates EU041867, EU041810<br />

Rhinocladiella phaeophora <strong>CBS</strong> 496.78*; IMI 287527 Soil Colombia EU041868, EU041811<br />

Rhinocladiella sp. <strong>CBS</strong> 264.49; MUCL 9904 Honey France EU041869, EU041812<br />

Rhodoveronaea varioseptata <strong>CBS</strong> 431.88* Bertia moriformis Germany EU041870, EU041813<br />

Thysanorea papuana <strong>CBS</strong> 212.96* — Papua New Guinea EU041871, EU041814<br />

Veronaea botryosa <strong>CBS</strong> 121.92 Xanthorrhoea preissii Australia EU041872, EU041815<br />

<strong>CBS</strong> 254.57*; IMI 070233; MUCL 9821 — Italy EU041873, EU041816<br />

<strong>CBS</strong> 350.65; IMI 115127; MUCL 7972 Goat dung India EU041874, EU041817<br />

Veronaea compacta <strong>CBS</strong> 268.75* — South Africa EU041876, EU041819<br />

Veronaea japonica <strong>CBS</strong> 776.83* On dead bamboo culm Japan EU041875, EU041818<br />

Veronaeopsis simplex <strong>CBS</strong> 588.66*; IMI 203547 Acacia karroo South Africa EU041877, EU041820<br />

Zasmidium cellare <strong>CBS</strong> 146.36 Wine cellar — EU041878, EU041821<br />

1<br />

ATCC: American Type Culture Collection, Virginia, U.S.A.; <strong>CBS</strong>: Centraalbureau voor Schimmelcultures, Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; CPC: Culture collection of Pedro Crous,<br />

housed at <strong>CBS</strong>; DAOM: Plant Research Institute, Department of Agriculture (Mycology), Ottawa, Canada; IFM: Research Center for Pathogenic Fungi <strong>and</strong> Microbial Toxicoses,<br />

Chiba University, Chiba, Japan; IMI: International Mycological Institute, CABI-Bioscience, U.K.; JCM: Japan Collection of Microorganism, RIKEN BioResource Center, Japan;<br />

MFC: Matsushima Fungus Collection, Kobe, Japan; MUCL: Mycotheque de l’ Université Catholique de Louvain, Louvain-la-Neuve, Belgium; NCPF: <strong>The</strong> National Collection of<br />

Pathogenic Fungi, Holborn, London, U.K.; OAC: Department of Botany <strong>and</strong> Genetics, University of Guelph, Ont., Canada; QM: Quartermaster Research <strong>and</strong> Developement<br />

Center, U.S. Army, MA, U.S.A.; UTMB: University of Texas Medical Branch, Texas, U.S.A.<br />

*Ex-type cultures.<br />

Results<br />

Phylogeny<br />

<strong>The</strong> manually adjusted alignment of the 28S rDNA data contained<br />

137 sequences (including the two outgroups) <strong>and</strong> 995 characters<br />

including alignment gaps. Of the 748 characters used in the<br />

phylogenetic analysis, 373 were parsimony-informative, 61 were<br />

variable <strong>and</strong> parsimony-uninformative, <strong>and</strong> 314 were constant.<br />

Neighbour-joining analysis using the three substitution models on<br />

the LSU alignment yielded trees with <strong>similar</strong> topology <strong>and</strong> bootstrap<br />

values. Parsimony analysis of the alignment yielded 5 000 equally<br />

most parsimonious trees, one of which is shown in Fig. 1 (TL =<br />

2 157, CI = 0.377, RI = 0.875, RC = 0.330). <strong>The</strong> Markov Chain<br />

Monte Carlo (MCMC) analysis of four chains started from a r<strong>and</strong>om<br />

tree topology <strong>and</strong> lasted 2 000 000 generations. Trees were saved<br />

each 1 000 generations, resulting in 2 000 trees. Burn-in was set<br />

at 500 000 generations after which the likelihood values were<br />

stationary, leaving 1 500 trees from which the consensus tree (Fig.<br />

2) <strong>and</strong> posterior probabilities (PP’s) were calculated. <strong>The</strong> average<br />

st<strong>and</strong>ard deviation of split frequencies was 0.043910 at the end<br />

of the run. Among the neighbour-joining, Bayesian <strong>and</strong> parsimony<br />

analyses, the trees differed in the hierarchical order of the main<br />

families <strong>and</strong> the support values (data not shown; e.g. the support<br />

within of the Capnodiales in Figs 1–2).<br />

<strong>The</strong> phylogenetic trees (Figs 1–2) show that the Ramichloridium<br />

species segregate into eight distinct clades, residing in the<br />

Capnodiales (Mycosphaerellaceae <strong>and</strong> Teratosphaeriaceae), the<br />

Chaetothyriales (Herpotrichiellaceae), the Pleosporales, <strong>and</strong> five<br />

other clades of which the relationships remain to be elucidated. <strong>The</strong><br />

type species of Ramichloridium, R. apiculatum, together with R.<br />

musae, R. cerophilum (Tubaki) de Hoog, R. indicum (Subram.) de<br />

Hoog, R. pini <strong>and</strong> three new species respectively isolated from Musa<br />

banksii, Strelitzia nicolai, <strong>and</strong> forest soil, reside in different parts of<br />

the Capnodiales clade (all in the Mycosphaerellaceae, except for the<br />

species from forest soil which clusters in the Teratosphaeriaceae).<br />

<strong>The</strong> second clade (in the Chaetothyriomycetes clade), including<br />

the human-pathogenic species R. mackenziei <strong>and</strong> R. basitonum,<br />

together with R. fasciculatum V. Rao & de Hoog <strong>and</strong> R. anceps<br />

(Sacc. & Ellis) de Hoog, groups together with Rhinocladiella in the<br />

Herpotrichiellaceae. <strong>The</strong> third clade (in the Sordariomycetes clade)<br />

includes R. obovoideum (Matsush.) de Hoog, which in a Blast<br />

search was found to have affinity with Carpoligna pleurothecii F.A.<br />

Fernández & Huhndorf (Chaetosphaeriales). <strong>The</strong> fourth clade (in<br />

the Sordariomycetes clade) includes a veronaea-like isolate from<br />

Bertia moriformis, with phylogenetic affinity to the Annulatascaceae<br />

(Sordariomycetidae). <strong>The</strong> fifth clade (in the Sordariomycetes clade)<br />

includes R. schulzeri var. schulzeri <strong>and</strong> R. schulzeri var. flexuosum<br />

de Hoog, the closest relatives being Thyridium vestitum (Fr.)<br />

Fuckel in the Thyridiaceae <strong>and</strong> Magnaporthe grisea (T.T. Hebert)<br />

M.E. Barr in the Magnaporthaceae. <strong>The</strong> sixth clade (in the Incertae<br />

sedis clade) includes R. subulatum de Hoog, R. epichloës (Ellis &<br />

Dearn.) de Hoog <strong>and</strong> a species isolated from the Poaceae. Three<br />

ramichloridium-like isolates from Rubus coreanus <strong>and</strong> Agrimonia<br />

pilosa form another unique clade (in the Incertae sedis clade)<br />

with uncertain affinity. Veronaea simplex Papendorf clusters as<br />

sister taxon to the Venturiaceae representing the eighth clade<br />

(Dothideomycetes). <strong>The</strong> type species of Periconiella, P. velutina,<br />

clusters within the Mycosphaerellaceae (Capnodiales clade),<br />

whereas P. papuana Aptroot resides in the Herpotrichiellaceae<br />

(Chaetothyriales clade). Veronaea botryosa Cif. & Montemart., the<br />

type species of Veronaea, also resides in the Herpotrichiellaceae.<br />

Taxonomy<br />

<strong>The</strong> species previously described in Ramichloridium share<br />

some morphological features, including erect, pigmented, more<br />

or less differentiated conidiophores, sympodially proliferating<br />

conidiogenous cells <strong>and</strong> predominantly aseptate conidia. Other than<br />

conidial morphology, features of the conidiogenous apparatus that<br />

60


Ramichloridium <strong>and</strong> allied genera<br />

appear to be more phylogenetically informative include pigmentation<br />

of vegetative hyphae, conidiophores <strong>and</strong> conidia, denticle density<br />

on the rachis, <strong>and</strong> structure of the scars. By integrating these data<br />

with the molecular data set, more natural genera are delineated,<br />

which are discussed below.<br />

Key to ramichloridium-like genera<br />

1. Conidiogenous cells integrated, terminal <strong>and</strong> lateral on creeping or ascending hyphae (differentiation between branched vegetative<br />

hyphae <strong>and</strong> conidiophores barely possible); conidiogenous loci bulging, more or less umbonate, apex rounded; occurring on rust<br />

pustules ......................................................................................................................................................................... Pseudovirgaria<br />

1. Conidiogenous cells integrated in distinct conidiophores; conidiogenous loci non-umbonate (flat, not prominent; subcylindrical or conical<br />

denticles; or terminally flat-tipped; or thickened <strong>and</strong> darkened); rarely occurring on rust pustules, but if so, with a raduliform rachis <strong>and</strong><br />

distinct denticles......................................................................................................................................................................................... 2<br />

2. Conidia 0–2(–3)-septate, conidial base truncate, retaining a marginal frill after liberation [anamorphs of Sordariomycetes]<br />

........................................................................................................................................................................................... Rhodoveronaea<br />

2. Conidial base without marginal frill ............................................................................................................................................................ 3<br />

3. Conidiophores composed of a well-developed erect stalk <strong>and</strong> a terminal branched head ........................................................................ 4<br />

3. Conidiophores unbranched or, if branched, branches loose, irregular or dichotomous, but not distinctly separated into stalk <strong>and</strong> branched<br />

head ........................................................................................................................................................................................................... 5<br />

4. Conidiophores dimorphic, either macronematous, dark brown with a dense apical cluster of branches or micronematous, undifferentiated,<br />

resembling vegetative hyphae; both kinds with a denticulate rachis; conidia predominantly 1-septate [anamorph of Chaetothyriales]<br />

.................................................................................................................................................................................................. Thysanorea<br />

4. Conidiophores monomorphic; branched head with fewer branches <strong>and</strong> looser; conidiogenous loci usually flat, non-prominent, less<br />

denticle-like; conidia aseptate to pluriseptate [anamorphs of Capnodiales] ............................................................................ Periconiella<br />

5. Rachis with denticles 1–1.5 µm long, denticles almost cylindrical; conidia at least partly in short chains ......................... Pleurothecium<br />

5. Rachis with denticles less than 1 µm long, denticles not cylindrical or denticles lacking, rachis with flat, barely prominent scars ........... 6<br />

6. Conidia predominantly septate .................................................................................................................................................................. 7<br />

6. Conidia predominantly aseptate ................................................................................................................................................................ 8<br />

7. Conidiophores up to 200 µm long; rachis straight, not to slightly geniculate; conidiogenous loci more or less flat, barely prominent,<br />

unthickened, slightly darkened [anamorphs of Chaetothyriales, Herpotrichiellaceae] ................................................................. Veronaea<br />

7. Conidiophores up to 60 µm long; rachis distinctly geniculate; conidiogenous loci denticle-like, prominent, up to 0.5 µm high, slightly<br />

thickened <strong>and</strong> darkened [anamorph of Pleosporales, Venturiaceae] ................................................................................... Veronaeopsis<br />

8. Vegetative mycelium entirely hyaline; rachis long, hyaline, with widely scattered pimple-shaped, terminally pointed, unpigmented<br />

denticles .............................................................................................................................................................................. Myrmecridium<br />

8. Vegetative mycelium at least partly pigmented; conidiogenous loci distinct, non-denticulate, somewhat darkened-refractive, or denticles,<br />

if present, neither pimple-shaped nor pointed ............................................................................................................................................ 9<br />

9. Rachis distinctly raduliform, with distinct, prominent blunt denticles, 0.5–1 µm long; scars <strong>and</strong> hila unthickened, but pigmented<br />

.................................................................................................................................................................................................. Radulidium<br />

9. Rachis not distinctly raduliform, at most subdenticulate; scars flat or only slightly prominent (subdenticulate), shorter ......................... 10<br />

10. Conidiophores usually poorly differentiated from the vegetative hyphae; conidial apparatus often loosely branched; exophiala-like<br />

budding cells usually present in culture [anamorphs of Chaetothyriales, Herpotrichiellaceae] .......................................... Rhinocladiella<br />

10. Conidiophores usually well differentiated from the vegetative mycelium (macronematous), usually unbranched; without exophiala-like<br />

states [anamorphs of Capnodiales] .................................................................................................................................. Ramichloridium<br />

Capnodiales (Mycosphaerellaceae, Teratosphaeriaceae)<br />

<strong>The</strong> type species of Ramichloridium, R. apiculatum, together<br />

with R. indicum cluster as a sister group to the Dissoconium de<br />

Hoog, Oorschot & Hijwegen clade in the Mycosphaerellaceae.<br />

Some other Ramichloridium species, including R. musae, R.<br />

biverticillatum Arzanlou & Crous, R. pini <strong>and</strong> R. cerophilum, are<br />

also allied with members of the Mycosphaerellaceae. Three<br />

additional new species are introduced for Ramichloridium isolates<br />

www.studiesinmycology.org<br />

from Musa banksii, Strelitzia nicolai, <strong>and</strong> forest soil. Periconiella<br />

velutina, the type species of Periconiella, which also resides in the<br />

Mycosphaerellaceae, is morphologically sufficiently distinct to retain<br />

its generic status. Two new species of Periconiella are introduced<br />

for isolates obtained from Turpinia pomifera <strong>and</strong> Ischyrolepis<br />

subverticellata in South Africa. Zasmidium cellare (Pers.) Fr., the<br />

type species of Zasmidium (Pers.) Fr., is also shown to cluster<br />

within the Mycosphaerellaceae.<br />

61


Arzanlou et al.<br />

10 changes<br />

Athelia epiphylla AY586633<br />

Paullicorticium ansatum AY586693<br />

100 Conioscyphascus varius AY484512<br />

100<br />

Conioscypha lignicola AY484513<br />

100 Carpoligna pleurothecii AF064645<br />

78<br />

Carpoligna pleurothecii AF064646<br />

Carpoligna pleurothecii AY544685<br />

98<br />

74 Pleurothecium obovoideum <strong>CBS</strong> 209.95<br />

Ascotaiwania hughesii AY316357<br />

100 Ascotaiwania hughesii AY094189<br />

Ophiostoma stenoceras DQ836904<br />

96<br />

Magnaporthe grisea AB026819<br />

100 Cryptadelphia polyseptata AY281102<br />

Cryptadelphia groenendalensis AY281103<br />

100<br />

100 82<br />

86<br />

96<br />

89<br />

Rhodoveronaea varioseptata <strong>CBS</strong> 431.88<br />

Annulatascus triseptatus AY780049<br />

Annulatascus triseptatus AY346257<br />

Thyridium vestitum AY544671<br />

Capronia pulcherrima AF050256<br />

Exophiala dermatitidis AF050270<br />

Capronia mansonii AY004338<br />

Rhinocladiella sp. <strong>CBS</strong> 264.49<br />

100<br />

100<br />

Myrmecridium schulzeri <strong>CBS</strong> 114996<br />

Myrmecridium flexuosum <strong>CBS</strong> 398.76<br />

Myrmecridium schulzeri <strong>CBS</strong> 188.96<br />

Myrmecridium schulzeri <strong>CBS</strong> 381.87<br />

Myrmecridium schulzeri <strong>CBS</strong> 305.73<br />

Myrmecridium schulzeri <strong>CBS</strong> 304.73<br />

Myrmecridium schulzeri <strong>CBS</strong> 100.54<br />

Myrmecridium schulzeri <strong>CBS</strong> 642.76<br />

Myrmecridium schulzeri <strong>CBS</strong> 134.68<br />

Myrmecridium schulzeri <strong>CBS</strong> 156.63<br />

Myrmecridium schulzeri <strong>CBS</strong> 325.74<br />

Rhinocladiella mackenziei <strong>CBS</strong> 368.92<br />

Rhinocladiella mackenziei AF050288<br />

Rhinocladiella mackenziei <strong>CBS</strong> 367.92<br />

Rhinocladiella mackenziei <strong>CBS</strong>102590<br />

Capronia coronata AF050242<br />

Rhinocladiella fasciculata <strong>CBS</strong> 132.86<br />

Rhinocladiella phaeophora <strong>CBS</strong> 496.78<br />

Rhinocladiella anceps AF050284<br />

Rhinocladiella anceps <strong>CBS</strong> 181.65<br />

Rhinocladiella anceps AF050285<br />

Rhinocladiella anceps <strong>CBS</strong> 157.54<br />

Fonsecaea pedrosoi AF050276<br />

Veronaea botryosa <strong>CBS</strong> 350.65<br />

Veronaea botryosa <strong>CBS</strong> 121.92<br />

Veronaea botryosa <strong>CBS</strong> 254.57<br />

Thysanorea papuana <strong>CBS</strong> 212.96<br />

Veronaea japonica <strong>CBS</strong> 776.83<br />

Veronaea compacta <strong>CBS</strong> 268.75<br />

99<br />

80 60<br />

100<br />

95<br />

88<br />

Rhinocladiella basitona <strong>CBS</strong> 101460<br />

Rhinocladiella atrovirens AF050289<br />

Exophiala jeanselmei AF050271<br />

Veronaeopsis simplex <strong>CBS</strong> 588.66<br />

Repetophragma goidanichii DQ408574<br />

Venturia pyrina EF114715<br />

Metacoleroa dickiei DQ384100<br />

Venturia chlorospora DQ384101<br />

Venturia inaequalis EF114713<br />

Venturia hanliniana AF050290<br />

54<br />

Venturia asperata EF114711<br />

Venturia carpophila AY849967<br />

Sordariomycetes<br />

Herpotrichiellaceae,<br />

Chaetothyriales,<br />

Chaetothyriomycetes<br />

Venturiaceae,<br />

Pleosporales,<br />

Dothideomycetes<br />

Fig. 1. (Page 62–63). One of 5 000 equally most parsimonious trees obtained from a heuristic search with simple taxon additions of the LSU sequence alignment using PAUP v.<br />

4.0b10. <strong>The</strong> scale bar shows 10 changes; bootstrap support values from 1 000 replicates are shown at the nodes. Thickened lines indicate the strict consensus branches <strong>and</strong> extype<br />

sequences are printed in bold face. <strong>The</strong> tree was rooted to two sequences obtained from GenBank (Athelia epiphylla AY586633 <strong>and</strong> Paullicorticium ansatum AY586693).<br />

Periconiella Sacc., in Sacc. & Berlese, Atti Ist. Veneto Sci., Ser. 6,<br />

3: 727. 1885.<br />

In vitro: Colonies with entire margin; aerial mycelium rather<br />

compact, raised, velvety, olivaceous-grey; reverse olivaceousblack.<br />

Submerged hyphae verrucose, hyaline, thin-walled, 1–3 µm<br />

wide; aerial hyphae subhyaline, later becoming dark brown, thickwalled,<br />

smooth. Conidiophores arising vertically from creeping<br />

hyphae, straight or flexuose, up to 260 µm long, dark brown at<br />

the base, paler towards the apex, thick-walled; in the upper part<br />

bearing short branches. Conidiogenous cells terminally integrated,<br />

polyblastic, smooth or verrucose, subcylindrical, mostly not or barely<br />

geniculate-sinuous, variable in length, subhyaline, later becoming<br />

pale brown, fertile part as wide as the basal part, proliferating<br />

sympodially, sometimes becoming septate <strong>and</strong> forming a short,<br />

straight rachis with pigmented, slightly thickened <strong>and</strong> hardly<br />

prominent, more or less flat scars. Conidia solitary, occasionally in<br />

short chains, 0–multi-septate, subhyaline to rather pale olivaceous<br />

or olivaceous-brown, smooth to verrucose, globose, ellipsoidal to<br />

obovoid or obclavate, with a slightly darkened <strong>and</strong> thickened hilum;<br />

conidial secession schizolytic.<br />

Type species: P. velutina (G. Winter) Sacc., Miscell. mycol. 2: 17.<br />

1884.<br />

62


Ramichloridium <strong>and</strong> allied genera<br />

54<br />

10 changes<br />

Fig. 1. (Continued).<br />

52<br />

76<br />

100 100<br />

98<br />

94<br />

71<br />

100<br />

77<br />

99<br />

90<br />

98<br />

Radulidium sp. <strong>CBS</strong> 115704<br />

Pseudovirgaria hyperparasitica CPC 10702<br />

Pseudovirgaria hyperparasitica CPC 10753<br />

Pseudovirgaria hyperparasitica CPC 10704<br />

96 Radulidium subulatum <strong>CBS</strong> 287.84<br />

99 Radulidium subulatum <strong>CBS</strong> 912.96<br />

Radulidium epichloës <strong>CBS</strong> 361.63<br />

78<br />

Radulidium epichloës AF050287<br />

64<br />

Radulidium subulatum <strong>CBS</strong> 405.76<br />

Radulidium subulatum <strong>CBS</strong> 101010<br />

Sporidesmium pachyanthicola DQ408557<br />

Staninwardia suttonii DQ923535<br />

Ramichloridium brasilianum <strong>CBS</strong> 283.92<br />

Batcheloromyces proteae <strong>CBS</strong> 110696<br />

Teratosphaeria alistairii DQ885901<br />

Teratosphaeria toledana DQ246230<br />

Readeriella considenianae DQ923527<br />

Teratosphaeria molleriana EU019292<br />

Teratosphaeria fibrillosa EU019282<br />

Catenulostroma macowanii EU019254<br />

67<br />

58 100<br />

100<br />

87<br />

100<br />

56<br />

52<br />

99 54<br />

99<br />

87<br />

Teratosphaeria suberosa DQ246235<br />

Cibiessia dimorphospora EU019258<br />

Teratosphaeria readeriellophora DQ246238<br />

Xenomeris juniperi EF114709<br />

Catenulostroma abietis EU019249<br />

Devriesia staurophora DQ008150<br />

Teratosphaeria parva DQ246240<br />

69<br />

78<br />

64<br />

Teratosphaeria parva DQ246243<br />

<strong>Cladosporium</strong> cladosporioides EU019262<br />

<strong>Cladosporium</strong> uredinicola EU019264<br />

<strong>Cladosporium</strong> bruhnei EU019261<br />

<strong>Cladosporium</strong> sphaerospermum EU019263<br />

Ramichloridium indicum <strong>CBS</strong> 171.96<br />

Dissoconium aciculare EU019266<br />

“Mycosphaerella” communis EU019267<br />

“Mycosphaerella” lateralis EU019268<br />

Ramichloridium apiculatum <strong>CBS</strong> 390.67<br />

Ramichloridium apiculatum <strong>CBS</strong> 400.76<br />

Ramichloridium apiculatum <strong>CBS</strong> 391.67<br />

Ramichloridium apiculatum <strong>CBS</strong> 156.59<br />

100<br />

100<br />

95<br />

99<br />

56<br />

55<br />

80<br />

Ramichloridium pini <strong>CBS</strong> 461.82<br />

Mycosphaerella endophytica DQ246252<br />

Mycosphaerella endophytica DQ246255<br />

Mycosphaerella gregaria DQ246251<br />

Mycosphaerella graminicola EU019297<br />

Septoria tritici EU019298<br />

Cercosporella centaureicola EU019257<br />

Mycosphaerella punctiformis AY490776<br />

Ramularia sp. EU019285<br />

Ramularia miae DQ885902<br />

Ramularia pratensis var. pratensis EU019284<br />

Mycosphaerella walkeri DQ267574<br />

Mycosphaerella parkii DQ246245<br />

Mycosphaerella madeirae DQ204756<br />

Periconiella levispora <strong>CBS</strong> 873.73<br />

Mycosphaerella marksii DQ246249<br />

Pseudocercospora epispermogonia DQ204758<br />

Pseudocercospora epispermogonia DQ204757<br />

Mycosphaerella intermedia DQ246248<br />

Mycosphaerella marksii DQ246250<br />

Periconiella arcuata <strong>CBS</strong> 113477<br />

Rasutoria pseudotsugae EF114704<br />

Rasutoria tsugae EF114705<br />

Periconiella velutina <strong>CBS</strong> 101950<br />

Periconiella velutina <strong>CBS</strong> 101948<br />

Periconiella velutina <strong>CBS</strong> 101949<br />

Ramichloridium biverticillatum <strong>CBS</strong> 335.36<br />

Ramichloridium musae <strong>CBS</strong> 365.36<br />

Ramichloridium musae <strong>CBS</strong> 190.63<br />

56<br />

56<br />

77<br />

66<br />

97<br />

Ramichloridium australiense <strong>CBS</strong> 121710<br />

Ramichloridium strelitziae <strong>CBS</strong> 121711<br />

Zasmidium cellare <strong>CBS</strong> 146.36<br />

Ramichloridium cerophilum <strong>CBS</strong> 103.59<br />

Ramichloridium cerophilum AF050286<br />

Incertae sedis<br />

Teratosphaeriaceae<br />

Davidiellaceae<br />

Mycosphaerellaceae<br />

Capnodiales, Dothideomycetes<br />

Notes: Periconiella is distinct from other ramichloridium-like genera<br />

by its conidiophores that are prominently branched in the upper<br />

part, <strong>and</strong> by its darkened, thickened conidial scars, that are more<br />

or less flat <strong>and</strong> non-prominent. Although conidiophores are also<br />

branched in the upper part in Thysanorea Arzanlou, W. Gams &<br />

Crous, the branching pattern in the latter <strong>genus</strong> is different from<br />

that of Periconiella. Thysanorea has a complex head consisting of<br />

up to six levels of branches, while in Periconiella the branching is<br />

limited, with mainly primary <strong>and</strong> secondary branches. Furthermore,<br />

Thysanorea is characterised by having dimorphic conidiophores<br />

<strong>and</strong> more or less prominent denticle-like conidiogenous loci.<br />

www.studiesinmycology.org<br />

63


Arzanlou et al.<br />

Athelia epiphylla AY586633<br />

Paullicorticium ansatum AY586693<br />

Veronaeopsis simplex <strong>CBS</strong> 588.66<br />

0.98 Repetophragma goidanichii DQ408574<br />

Metacoleroa dickiei DQ384100<br />

0.79 Venturia pyrina EF114715<br />

1.00 Venturia hanliniana AF050290<br />

Venturia inaequalis EF114713<br />

0.84 Venturia chlorospora DQ384101<br />

Venturia asperata EF114711<br />

1.00<br />

0.1 expected changes per site<br />

0.99<br />

0.52<br />

0.74<br />

1.00<br />

1.00<br />

0.97<br />

1.00<br />

0.80<br />

0.63<br />

0.99<br />

0.50<br />

0.68<br />

0.69<br />

1.00<br />

0.58<br />

Venturia carpophila AY849967<br />

Capronia pulcherrima AF050256<br />

Capronia coronata AF050242<br />

Rhinocladiella phaeophora <strong>CBS</strong> 496.78<br />

Rhinocladiella fasciculata <strong>CBS</strong> 132.86<br />

1.00<br />

Exophiala dermatitidis AF050270<br />

Capronia mansonii AY004338<br />

Rhinocladiella mackenziei <strong>CBS</strong> 368.92<br />

Rhinocladiella mackenziei AF050288<br />

Rhinocladiella mackenziei <strong>CBS</strong> 367.92<br />

Rhinocladiella mackenziei <strong>CBS</strong> 102590<br />

1.00<br />

Rhinocladiella anceps AF050284<br />

Rhinocladiella anceps <strong>CBS</strong> 181.65<br />

Rhinocladiella anceps AF050285<br />

Rhinocladiella anceps <strong>CBS</strong> 157.54<br />

Rhinocladiella sp. <strong>CBS</strong> 264.49<br />

Exophiala jeanselmei AF050271<br />

Rhinocladiella atrovirens AF050289<br />

Rhinocladiella basitona <strong>CBS</strong> 101460<br />

Fonsecaea pedrosoi AF050276<br />

Thysanorea papuana <strong>CBS</strong> 212.96<br />

Veronaea japonica <strong>CBS</strong> 776.83<br />

Veronaea compacta <strong>CBS</strong> 268.75<br />

Veronaea botryosa <strong>CBS</strong> 350.65<br />

Veronaea botryosa <strong>CBS</strong> 121.92<br />

Veronaea botryosa <strong>CBS</strong> 254.57<br />

0.84<br />

0.99<br />

0.96<br />

0.99<br />

0.91<br />

1.00<br />

1.00 Conioscyphascus varius AY484512<br />

Conioscypha lignicola AY484513<br />

1.00 Carpoligna pleurothecii AF064646<br />

1.00<br />

Carpoligna pleurothecii AF064645<br />

Carpoligna pleurothecii AY544685<br />

0.99 1.00 Pleurothecium obovoideum <strong>CBS</strong> 209.95<br />

1.00 Ascotaiwania hughesii AY316357<br />

Ascotaiwania hughesii AY094189<br />

Magnaporthe grisea AB026819<br />

Ophiostoma stenoceras DQ836904<br />

Cryptadelphia polyseptata AY281102<br />

Cryptadelphia groenendalensis AY281103<br />

1.00<br />

1.00<br />

1.00<br />

0.61<br />

0.87<br />

0.71<br />

0.92<br />

1.00<br />

Venturiaceae,<br />

Pleosporales,<br />

Dothideomycetes<br />

Rhodoveronaea varioseptata <strong>CBS</strong> 431.88<br />

Annulatascus triseptatus AY780049<br />

Annulatascus triseptatus AY346257<br />

Thyridium vestitum AY544671<br />

Myrmecridium schulzeri <strong>CBS</strong> 114996<br />

Myrmecridium schulzeri <strong>CBS</strong> 188.96<br />

Myrmecridium schulzeri <strong>CBS</strong> 381.87<br />

Myrmecridium schulzeri <strong>CBS</strong> 305.73<br />

Myrmecridium schulzeri <strong>CBS</strong> 304.73<br />

Myrmecridium flexuosum <strong>CBS</strong> 398.76<br />

Myrmecridium schulzeri <strong>CBS</strong> 100.54<br />

Myrmecridium schulzeri <strong>CBS</strong> 642.76<br />

Myrmecridium schulzeri <strong>CBS</strong> 134.68<br />

Myrmecridium schulzeri <strong>CBS</strong> 156.63<br />

Myrmecridium schulzeri <strong>CBS</strong> 325.74<br />

Herpotrichiellaceae,<br />

Chaetothyriales,<br />

Chaetothyriomycetes<br />

Sordariomycetes<br />

Fig. 2. (Page 64–65). Consensus phylogram (50 % majority rule) of 1 500 trees resulting from a Bayesian analysis of the LSU sequence alignment using MrBayes v. 3.1.2.<br />

Bayesian posterior probabilities are indicated at the nodes. Ex-type sequences are printed in bold face. <strong>The</strong> tree was rooted to two sequences obtained from GenBank (Athelia<br />

epiphylla AY586633 <strong>and</strong> Paullicorticium ansatum AY586693).<br />

Periconiella velutina (G. Winter) Sacc., Miscell. mycol. 2: 17.<br />

1884. Fig. 3.<br />

Basionym: Periconia velutina G. Winter, Hedwigia 23: 174. 1884.<br />

In vitro: Submerged hyphae verrucose, hyaline, thin-walled, 1–3<br />

µm wide; aerial hyphae subhyaline, later becoming dark brown,<br />

thick-walled, smooth. Conidiophores arising vertically from creeping<br />

hyphae, straight or flexuose, up to 260 µm long, dark brown at the<br />

base, paler towards the apex, thick-walled; in the upper part bearing<br />

short branches, 10–35 µm long. Conidiogenous cells mostly<br />

terminally integrated, sometimes discrete, smooth or verrucose,<br />

cylindrical, variable in length, subhyaline, later becoming pale<br />

brown, fertile part as wide as the basal part, proliferating sympodially,<br />

sometimes becoming septate <strong>and</strong> forming a short, straight rachis<br />

with pigmented, slightly thickened <strong>and</strong> hardly prominent, more<br />

or less flat scars, less than 1 µm diam. Conidia 0(–1)-septate,<br />

subhyaline, thin-walled, verrucose or smooth, globose, ellipsoidal<br />

to obovoid, (7–)8–9(–11) × (2.5–)3(–4) µm, with a slightly darkened<br />

<strong>and</strong> thickened hilum, 1.5–2 µm diam.<br />

Cultural characteristics: Colonies on MEA slow-growing, reaching<br />

4 mm diam after 14 d at 24 °C, with entire margin; aerial mycelium<br />

rather compact, raised, velvety, olivaceous-grey; reverse<br />

olivaceous-black.<br />

Specimens examined: South Africa, Cape Town, on Brabejum stellatifolium, P.<br />

MacOwan, herb. G. Winter (B), lectotype selected here; Cape Town, on leaves<br />

of Brabejum stellatifolium (= B. stellatum), P. Mac-Owan, PAD, F42165, F462166,<br />

isolectotypes; Stellenbosch, Jonkershoek Nature Reserve, on Brabejum<br />

stellatifolium, 21 Jan. 1999, J.E. Taylor, epitype designated here <strong>CBS</strong> H-15612,<br />

cultures ex-epitype <strong>CBS</strong> 101948–101950.<br />

64


Ramichloridium <strong>and</strong> allied genera<br />

0.1 expected changes per site<br />

Fig. 2. (Continued).<br />

0.64<br />

0.61<br />

0.96<br />

0.99<br />

0.87<br />

1.00<br />

1.00<br />

1.00<br />

0.68<br />

0.99<br />

1.00<br />

1.00<br />

1.00<br />

1.00<br />

0.99<br />

0.86<br />

Pseudovirgaria hyperparasitica CPC 10702<br />

Pseudovirgaria hyperparasitica CPC 10753<br />

Pseudovirgaria hyperparasitica CPC 10704<br />

Radulidium sp. <strong>CBS</strong> 115704<br />

0.95 Radulidium subulatum <strong>CBS</strong> 287.84<br />

Radulidium subulatum <strong>CBS</strong> 912.96<br />

Radulidium epichloës <strong>CBS</strong> 361.63<br />

Radulidium epichloës AF050287<br />

Radulidium subulatum <strong>CBS</strong> 405.76<br />

Radulidium subulatum <strong>CBS</strong> 101010<br />

0.60<br />

0.57<br />

<strong>Cladosporium</strong> cladosporioides EU019262<br />

0.91<br />

<strong>Cladosporium</strong> uredinicola EU019264<br />

1.00 <strong>Cladosporium</strong> bruhnei EU019261<br />

0.81 <strong>Cladosporium</strong> sphaerospermum EU019263<br />

1.00 Sporidesmium pachyanthicola DQ408557<br />

Staninwardia suttonii DQ923535<br />

1.00 Ramichloridium brasilianum <strong>CBS</strong> 283.92<br />

1.00 Batcheloromyces proteae EU019247<br />

Teratosphaeria alistairii DQ885901<br />

Readeriella considenianae DQ923527<br />

Teratosphaeria toledana DQ246230<br />

Teratosphaeria molleriana EU019292<br />

Teratosphaeria fibrillosa EU019282<br />

Catenulostroma macowanii EU019254<br />

0.63<br />

0.73<br />

0.72<br />

0.95<br />

1.00<br />

1.00<br />

1.00<br />

1.00<br />

0.64<br />

1.00<br />

1.00<br />

Teratosphaeria suberosa DQ246235<br />

Cibiessia dimorphospora EU019258<br />

Teratosphaeria readeriellophora DQ246238<br />

Teratosphaeria parva DQ246240<br />

Teratosphaeria parva DQ246243<br />

Devriesia staurophora DQ008150<br />

Xenomeris juniperi EF114709<br />

Catenulostroma abietis EU019249<br />

Ramichloridium indicum <strong>CBS</strong> 171.96<br />

Dissoconium aciculare EU019266<br />

“Mycosphaerella” communis EU019267<br />

“Mycosphaerella” lateralis EU019268<br />

Ramichloridium apiculatum <strong>CBS</strong> 390.67<br />

Ramichloridium apiculatum <strong>CBS</strong> 400.76<br />

Ramichloridium apiculatum <strong>CBS</strong> 391.67<br />

Ramichloridium apiculatum <strong>CBS</strong> 156.59<br />

0.99<br />

0.72<br />

1.00<br />

Mycosphaerella endophytica DQ246252<br />

Mycosphaerella endophytica DQ246255<br />

Mycosphaerella gregaria DQ246251<br />

Cercosporella centaureicola EU019257<br />

Mycosphaerella punctiformis AY490776<br />

Ramularia sp. EU019285<br />

Ramularia miae DQ885902<br />

Ramularia pratensis var. pratensis EU019284<br />

Mycosphaerella graminicola EU019297<br />

Septoria tritici EU019298<br />

Ramichloridium pini <strong>CBS</strong> 461.82<br />

1.00 Mycosphaerella walkeri DQ267574<br />

Mycosphaerella parkii DQ246245<br />

Mycosphaerella madeirae DQ204756<br />

0.95 Periconiella levispora <strong>CBS</strong> 873.73<br />

0.76<br />

0.60<br />

0.97<br />

0.64<br />

0.80<br />

1.00<br />

0.90<br />

0.95<br />

0.97<br />

1.00<br />

0.64<br />

0.99<br />

0.94<br />

0.97<br />

0.57<br />

Mycosphaerella marksii DQ246249<br />

Pseudocercospora epispermogonia DQ204758<br />

Pseudocercospora epispermogonia DQ204757<br />

Mycosphaerella intermedia DQ246248<br />

Mycosphaerella marksii DQ246250<br />

Periconiella arcuata <strong>CBS</strong> 113477<br />

Rasutoria pseudotsugae EF114704<br />

Rasutoria tsugae EF114705<br />

Periconiella velutina <strong>CBS</strong> 101950<br />

Periconiella velutina <strong>CBS</strong> 101948<br />

Periconiella velutina <strong>CBS</strong> 101949<br />

0.96<br />

0.97<br />

0.97<br />

0.96<br />

Ramichloridium biverticillatum <strong>CBS</strong> 335.36<br />

Ramichloridium musae <strong>CBS</strong> 365.36<br />

Ramichloridium musae <strong>CBS</strong> 190.63<br />

Ramichloridium australiense <strong>CBS</strong> 121710<br />

Ramichloridium strelitziae <strong>CBS</strong> 121711<br />

Zasmidium cellare <strong>CBS</strong> 146.36<br />

Ramichloridium cerophilum <strong>CBS</strong> 103.59<br />

Ramichloridium cerophilum AF050286<br />

Incertae sedis<br />

Davidiellaceae<br />

Teratosphaeriaceae<br />

Mycosphaerellaceae<br />

Capnodiales, Dothideomycetes<br />

Periconiella arcuata Arzanlou, S. Lee & Crous, sp. nov. MycoBank<br />

MB504547. Figs 4, 7A.<br />

Etymology: Named after its curved conidia.<br />

Ab aliis speciebus Periconiellae conidiis obclavatis, rectis vel curvatis, (30–)53–61(–<br />

79) × (3–)5(–7) µm, distinguenda.<br />

Submerged hyphae smooth, hyaline, thin-walled, 2 µm wide;<br />

aerial hyphae pale brown, smooth or verrucose, slightly narrower.<br />

Conidiophores arising vertically from creeping hyphae, straight<br />

or flexuose, up to 300 µm long, dark brown at the base, paler<br />

www.studiesinmycology.org<br />

towards the apex, thick-walled; loosely branched in the upper part,<br />

bearing short branches. Conidiogenous cells integrated, cylindrical,<br />

variable in length, 20–50 µm long, subhyaline, later becoming<br />

pale brown, fertile part as wide as the basal part, proliferating<br />

sympodially, forming a geniculate conidium-bearing rachis with<br />

pigmented <strong>and</strong> thickened, prominent, cone-shaped scars, 1 µm<br />

diam. Conidia formed singly, obclavate, straight or mostly curved,<br />

0(–4)-septate, coarsely verrucose, pale olive, thin-walled, tapering<br />

towards the apex, (30–)53–61(–79) × (3–)5(–7) µm, with a narrowly<br />

truncate base <strong>and</strong> a darkened, hardly thickened hilum, 2 µm diam;<br />

microcyclic conidiation observed in culture.<br />

65


Arzanlou et al.<br />

Fig. 3. Periconiella velutina (<strong>CBS</strong> 101948). A–B. Macronematous conidiophores with short branches in the upper part. C. Sympodially proliferating conidiogenous cell with<br />

darkened <strong>and</strong> slightly thickened scars. D. Conidia. Scale bar = 10 µm.<br />

Fig. 4. Periconiella arcuata (<strong>CBS</strong> 113477). A–B. Sympodially proliferating conidiogenous cells with darkened, thickened <strong>and</strong> cone-shaped scars. C–E. Macronematous<br />

conidiophores with loose branches in the upper part. F–I. Conidia. Scale bar = 10 µm.<br />

66


Ramichloridium <strong>and</strong> allied genera<br />

Fig. 5. Periconiella levispora (<strong>CBS</strong> 873.73). A–C. Conidial apparatus at different stages of development, which gives rise to macronematous conidiophores with dense branches<br />

in the upper part. D. Sympodially proliferating conidiogenous cells with darkened <strong>and</strong> somewhat protruding scars. E–F. Conidia with truncate base <strong>and</strong> darkened hilum. Scale<br />

bar = 10 µm.<br />

Fig. 6. A. Pseudovirgaria hyperparasitica (<strong>CBS</strong> 121739 = CPC 10753). B. Periconiella levispora (<strong>CBS</strong> 873.73). Scale bar = 10 µm.<br />

www.studiesinmycology.org<br />

67


Arzanlou et al.<br />

Cultural characteristics: Colonies on MEA reaching 12 mm diam<br />

after 14 d at 24 °C, with entire, smooth, sharp margin; mycelium<br />

compacted, becoming hairy, colonies up to 1 mm high; surface<br />

olivaceous to olivaceous-grey, reverse dark grey-olivaceous to<br />

olivaceous-black.<br />

Specimen examined: South Africa, Western Cape Province, Kogelberg, on dead<br />

culms of Ischyrolepis subverticillata, May 2001, S. Lee, holotype <strong>CBS</strong> H-19927,<br />

culture ex-type <strong>CBS</strong> 113477.<br />

Periconiella levispora Arzanlou, W. Gams & Crous, sp. nov.<br />

MycoBank MB504546. Figs 5–6B.<br />

Etymology: (Latin) levis = smooth.<br />

A simili Periconiella velutina conidiis levibus et maioribus, ad 23 μm longis<br />

distinguenda.<br />

In vitro: Submerged hyphae smooth, hyaline, thin-walled, 2–2.5 µm<br />

wide; aerial hyphae subhyaline, later becoming dark brown, thickwalled,<br />

smooth. Conidiophores arising vertically from creeping<br />

aerial hyphae, dark brown at the base, paler towards the apex,<br />

thick-walled; in the upper part bearing several short branches,<br />

up to 120 µm long. Conidiogenous cells integrated, occasionally<br />

discrete, cylindrical, variable in length, 10–20 µm long, subhyaline,<br />

later becoming pale brown, fertile part as wide as the basal part,<br />

proliferating sympodially, forming a short rachis with pigmented<br />

<strong>and</strong> slightly thickened, somewhat protruding scars, less than 1<br />

µm diam. Conidia solitary, 0(–2)-septate, smooth, pale olivaceous,<br />

cylindrical, ellipsoidal, pyriform to clavate, (7–)11–14(–23) × (3–)4–<br />

5(–6) µm, with a truncate base <strong>and</strong> a darkened, slightly thickened<br />

hilum, 2 µm diam.<br />

Cultural characteristics: Colonies on MEA slow-growing, reaching<br />

5 mm diam after 14 d at 24 °C, with entire margin; aerial mycelium<br />

compact, raised, velvety, olivaceous-grey; reverse olivaceousblack.<br />

Basionym: Chloridium apiculatum J.H. Mill., Giddens & A.A. Foster,<br />

Mycologia 49: 789. 1957.<br />

≡ Veronaea apiculata (J.H. Mill., Giddens & A.A. Foster) M.B. Ellis, in Ellis,<br />

More Dematiaceous Hyphomycetes: 209. 1976.<br />

[non Rhinocladiella apiculata Matsush., in Matsushima, Icon.<br />

Microfung. Mats. lect.: 122. 1975].<br />

= Rhinocladiella indica Agarwal, Lloydia 32: 388. 1969.<br />

[non Chloridium indicum Subram., Proc. Indian Acad. Sci., Sect. B,<br />

42: 286. 1955].<br />

In vitro: Submerged hyphae hyaline to subhyaline, thin-walled,<br />

1–2.5 µm wide; aerial hyphae slightly darker, smooth-walled.<br />

Conidiophores generally arising at right angles from creeping aerial<br />

hyphae, straight, unbranched, thick-walled, dark brown, continuous<br />

or with 1–2(–3) additional thin septa, up to 100 µm long; intercalary<br />

cells 10–28 µm long. Conidiogenous cells integrated, terminal,<br />

smooth, thick-walled, golden-brown, straight, cylindrical, 25–37(–<br />

47) × 2–3.5 µm; proliferating sympodially, resulting in a straight<br />

rachis with conspicuous conidiogenous loci; scars prominent,<br />

crowded, slightly pigmented, less than 1 µm diam. Conidia solitary,<br />

obovate to obconical, pale brown, finely verrucose, (3–)5–5.5(–7.5)<br />

× (2–)2.5–3(–4) µm, hilum conspicuous, slightly pigmented, about<br />

1 µm diam.<br />

Cultural characteristics: Colonies on MEA reaching 35 mm diam<br />

after 14 d at 24 °C; minimum temperature for growth above 6 °C,<br />

optimum 24 °C, maximum 30 °C. Colonies raised, velvety, dense,<br />

with entire margin; surface olivaceous-green, reverse olivaceousblack,<br />

often with a diffusing citron-yellow pigment.<br />

Specimens examined: Pakistan, Lahore, from soil, A. Kamal, <strong>CBS</strong> 400.76 = IMI<br />

088021. South Africa, from preserved Cucumis sativus in 8-oxyquinoline sulphate,<br />

M.C. Papendorf, <strong>CBS</strong> 390.67; Potchefstroom, from Aloe sp., M.C. Papendorf, <strong>CBS</strong><br />

391.67. U.S.A., Georgia, isolated from forest soil, <strong>CBS</strong> 156.59 = ATCC 13211 = IMI<br />

100716 = QM 7716, ex-type culture.<br />

Specimen examined: Sri Lanka, Hakgala Botanic Gardens, on dead leaves of<br />

Turpinia pomifera, Jan. 1973, W. Gams, holotype <strong>CBS</strong> H-15611, culture ex-type<br />

<strong>CBS</strong> 873.73.<br />

Ramichloridium Stahel ex de Hoog, Stud. Mycol. 15: 59. 1977.<br />

In vitro: Colonies flat to raised, with entire margin; surface<br />

olivaceous-green to olivaceous-black. Mycelium consisting of<br />

submerged <strong>and</strong> aerial hyphae; submerged hyphae hyaline to<br />

subhyaline, thin-walled, aerial hyphae smooth or verrucose.<br />

Conidiophores straight, unbranched, rarely branched, thick-walled,<br />

dark brown (darker than the subtending hyphae), continuous or<br />

with several additional thin septa. Conidiogenous cells integrated,<br />

terminal, polyblastic, smooth, thick-walled, golden-brown, apical<br />

part subhyaline, with sympodial proliferation, straight or flexuose,<br />

geniculate or nodose, with conspicuous conidiogenous loci;<br />

scars crowded or scattered, unthickened, unpigmented to faintly<br />

pigmented, or slightly prominent denticles. Conidia solitary, 0–1-<br />

septate, subhyaline to pale brown, smooth to coarsely verrucose,<br />

rather thin-walled, obovate, obconical or globose to ellipsoidal,<br />

fusiform, with a somewhat prominent, slightly pigmented hilum;<br />

conidial secession schizolytic.<br />

Type species: R. apiculatum (J.H. Mill., Giddens & A.A. Foster) de<br />

Hoog, Stud. Mycol. 15: 69. 1977.<br />

Ramichloridium apiculatum (J.H. Mill., Giddens & A.A. Foster) de<br />

Hoog, Stud. Mycol. 15: 69. 1977. Fig. 8.<br />

Fig. 7. A. Periconiella arcuata (<strong>CBS</strong> 113477). B. Myrmecridium schulzeri (<strong>CBS</strong><br />

325.74). C. Thysanorea papuana (<strong>CBS</strong> 212.96). Scale bars = 10 µm.<br />

68


Ramichloridium <strong>and</strong> allied genera<br />

Fig. 8. Ramichloridium apiculatum (<strong>CBS</strong> 156.59). A–C. Macronematous conidiophores with sympodially proliferating conidiogenous cells, which give rise to a conidium-bearing<br />

rachis with crowded <strong>and</strong> prominent scars. D. Conidia. Scale bar = 10 µm.<br />

Fig. 9. Ramichloridium australiense (<strong>CBS</strong> 121710). A–C. Macronematous conidiophores with thick-walled <strong>and</strong> warted subtending hyphae. D. Sympodially proliferating<br />

conidiogenous cell, which results in a short rachis with darkened <strong>and</strong> slightly thickened scars. E. Conidia. Scale bar = 10 µm.<br />

Ramichloridium australiense Arzanlou & Crous, sp. nov.<br />

MycoBank MB504548. Figs 9–10A.<br />

Etymology: Named after its country of origin, Australia.<br />

Ab aliis speciebus Ramichloridii conidiophoris ex hyphis verrucosis, crassitunicatis<br />

ortis distinguendum.<br />

In vitro: Submerged hyphae hyaline, smooth, thin-walled, 1–2 µm<br />

wide; aerial hyphae pale brown, warted. Conidiophores arising<br />

vertically <strong>and</strong> clearly differentiated from creeping aerial hyphae, up<br />

to 400 µm tall, with several additional thin septa; intercalary cells,<br />

8–40 × 2–5 µm, from the broadest part at the base tapering towards<br />

the apex, subhyaline, later becoming pale brown <strong>and</strong> warted in the<br />

lower part. Subtending hyphae thick-walled, warted. Conidiogenous<br />

cells integrated, terminal, 10–18 µm long, proliferating sympodially,<br />

www.studiesinmycology.org<br />

giving rise to a short rachis with conspicuous conidiogenous loci;<br />

scars slightly thickened <strong>and</strong> darkened, about 1 µm diam. Conidia<br />

solitary, aseptate, thin-walled, smooth, subhyaline, subcylindrical<br />

to obclavate, (10–)12–15(–23) × 2.5–3 µm, with a truncate base<br />

<strong>and</strong> a slightly darkened <strong>and</strong> thickened hilum,1.5–2 µm diam, rarely<br />

fusing at the basal part.<br />

Cultural characteristics: Colonies on MEA rather slow growing,<br />

reaching 8 mm diam after 14 d at 24 °C, with entire, smooth<br />

margin; mycelium flat, olivaceous-grey, becoming granular, with<br />

gelatinous droplets at the margin developing with aging; reverse<br />

pale olivaceous-grey.<br />

Specimen examined: Australia, Queensl<strong>and</strong>, Mount Lewis, Mount Lewis Road,<br />

16°34’47.2” S, 145°19’7” E, 538 m alt., on Musa banksii leaf, Aug. 2006, P.W. Crous<br />

<strong>and</strong> B. Summerell, holotype <strong>CBS</strong> H-19928, culture ex-type <strong>CBS</strong> 121710.<br />

69


Arzanlou et al.<br />

Fig. 10. A. Ramichloridium australiense (<strong>CBS</strong> 121710). B. Ramichloridium brasilianum (<strong>CBS</strong> 283.92). C. Radulidium subulatum (<strong>CBS</strong> 405.76). D. Rhodoveronaea varioseptata<br />

(<strong>CBS</strong> 431.88). Scale bar = 10 µm.<br />

Fig. 11. Ramichloridium musae (<strong>CBS</strong> 365.36). A. Conidiophores with loose branches. B–D. Sympodially proliferating conidiogenous cells, resulting in a long conidium-bearing<br />

rachis. E. Rachis with hardly prominent, slightly darkened scars. F. Conidia. Scale bars = 10 µm.<br />

70


Ramichloridium <strong>and</strong> allied genera<br />

Fig. 12. Ramichloridium biverticillatum (<strong>CBS</strong> 335.36). A–B. Profusely branched <strong>and</strong> biverticillate conidiophores. C. Sympodially proliferating conidiogenous cells,<br />

which give rise to a conidium-bearing rachis with crowded, slightly pigmented <strong>and</strong> thickened scars. D. Conidia. Scale bar = 10 µm.<br />

Fig. 13. Ramichloridium brasilianum (<strong>CBS</strong> 283.92). A–B. Macronematous conidiophores with sympodially proliferating conidiogenous cells, resulting in a conidiumbearing<br />

rachis. C. Rachis with crowded <strong>and</strong> slightly pigmented scars. D. Conidia. Scale bar = 10 µm.<br />

Fig. 14. Ramichloridium cerophilum (<strong>CBS</strong> 103.59). A–C. Conidial apparatus at different stages of development, resulting in macronematous conidiophores <strong>and</strong><br />

sympodially proliferating conidiogenous cells. D–E. Formation of secondary conidia. F. Conidia. Scale bar = 10 µm.<br />

www.studiesinmycology.org<br />

71


Arzanlou et al.<br />

Ramichloridium musae (Stahel ex M.B. Ellis) de Hoog, Stud.<br />

Mycol. 15: 62. 1977. Fig. 11.<br />

Basionym: Veronaea musae Stahel ex M.B. Ellis, in Ellis, More<br />

Dematiaceous Hyphomycetes: 209. 1976.<br />

≡ Chloridium musae Stahel, Trop. Agric., Trinidad 14: 43. 1937 (nom.<br />

inval. Art. 36).<br />

Misapplied name: Chloridium indicum Subram., sensu Batista &<br />

Vital, Anais Soc. Biol. Pernambuco 15: 379. 1957.<br />

In vitro: Submerged hyphae smooth, hyaline, thin-walled, 1–2 µm<br />

wide; aerial hyphae subhyaline, smooth. Conidiophores arising<br />

vertically <strong>and</strong> mostly sharply differentiated from creeping aerial<br />

hyphae, golden-brown; unbranched, rarely branched in the upper<br />

part, up to 250 µm tall, with up to 6 additional thin septa, cells<br />

23–40 × 2–2.5 µm, basal cell occasionally inflated. Conidiogenous<br />

cells terminally integrated, cylindrical, variable in length, 10–40 µm<br />

long, golden-brown near the base, subhyaline to pale brown near<br />

the end, fertile part as wide as the basal part, later also becoming<br />

septate; rachis elongating sympodially, 2–2.5 µm wide, with hardly<br />

prominent, scattered, slightly pigmented scars, about 0.5 µm diam.<br />

Conidia solitary, aseptate, hyaline to subhyaline, ellipsoidal, (4–)7–<br />

8(–12) × 2–3 µm, smooth or verruculose, thin-walled, with slightly<br />

darkened hilum, about 1 µm diam.<br />

Cultural characteristics: Colonies on MEA slow-growing, reaching<br />

27 mm diam after 14 d at 24 °C, with entire, smooth, sharp margin;<br />

mycelium mostly submerged, some floccose to lanose aerial<br />

mycelium in the olivaceous-grey centre, becoming pale pinkish<br />

olivaceous towards the margin; reverse pale orange.<br />

Specimens examined: Cameroon, from Musa sapientum, J.E. Heron, <strong>CBS</strong> 169.61<br />

= ATCC 15681 = IMI 079492 = DAOM 84655 = MUCL 2689; from Musa sapientum,<br />

J. Brun, <strong>CBS</strong> 190.63 = MUCL 9557. Surinam, Paramaribo, from Musa sapientum<br />

leaf, G. Stahel, <strong>CBS</strong> 365.36 = JCM 6973 = MUCL 9556, ex-type strain of Chloridium<br />

musae; from Musa sapientum, G. Stahel, <strong>CBS</strong> 365.36; dried culture preserved as<br />

<strong>CBS</strong> H-19933.<br />

Ramichloridium biverticillatum Arzanlou & Crous, nom. nov.<br />

MycoBank MB504549. Fig. 12.<br />

Basionym: Periconiella musae Stahel ex M.B. Ellis, Mycol. Pap.<br />

111: 5. 1967.<br />

[non Ramichloridium musae (Stahel ex M.B. Ellis) de Hoog, 1977].<br />

≡ Ramichloridium musae Stahel, Trop. Agric., Trinidad 14: 43. 1937 (nom.<br />

inval. Art. 36).<br />

= Ramichloridium musae (Stahel ex M.B. Ellis) de Hoog, Stud. Mycol. 15: 62.<br />

1977, sensu de Hoog, p.p.<br />

Etymology: Named after its biverticillate conidiophores.<br />

In vitro: Submerged hyphae smooth, hyaline, thin-walled, 1–2<br />

µm wide; aerial hyphae subhyaline, smooth, slightly darker.<br />

Conidiophores arising vertically from creeping aerial hyphae, pale<br />

brown, profusely branched, biverticillate, with up to three levels<br />

of main branches; branches tapering distally, 2–3 µm wide at the<br />

base, approx. 2 µm wide in the upper part, up to 250 µm long.<br />

Conidiogenous cells terminally integrated, cylindrical, variable in<br />

length, 15–50 µm long, rachis straight or geniculate, pale brown,<br />

as wide as the basal part; elongating sympodially, forming a rachis<br />

with crowded, slightly darkened <strong>and</strong> thickened minute scars, less<br />

than 0.5 µm wide. Conidia solitary, aseptate, hyaline to subhyaline,<br />

dacryoid to pyriform, (2–)3–4(–6) × (1.5–)2(–2.5) µm, smooth, thinwalled,<br />

with an inconspicuous hilum.<br />

Cultural characteristics: Colonies on MEA slow-growing, reaching<br />

16 mm diam after 14 d at 24 °C, with entire, smooth, sharp margin,<br />

rather compact, velvety; surface vinaceous-buff to olivaceous-buff;<br />

reverse buff.<br />

Specimen examined: Surinam, from Musa sapientum, Aug. 1936, G. Stahel, <strong>CBS</strong><br />

335.36.<br />

Notes: Ramichloridium biverticillatum is a new name based on<br />

Periconiella musae. <strong>The</strong> species is distinct from R. musae because<br />

of its profusely branched conidiophores, <strong>and</strong> conidia that are smaller<br />

(2–5 × 1.5–2.5 µm) than those of R. musae (5–11 × 2–3 µm).<br />

Ramichloridium brasilianum Arzanlou & Crous, sp. nov.<br />

MycoBank MB504550. Figs 10B, 13.<br />

Etymology: Named after its country of origin, Brazil.<br />

A simili Ramichloridio cerophilo conidiis minoribus, ad 8 μm longis, et conidiis<br />

secundariis absentibus distinguendum.<br />

In vitro: Submerged hyphae pale olivaceous, smooth or slightly<br />

rough, 1.5–2 µm wide; aerial hyphae olivaceous, smooth or rough,<br />

narrower <strong>and</strong> darker than the submerged hyphae. Conidiophores<br />

unbranched, arising vertically from creeping aerial hyphae, straight<br />

or flexuose, dark brown, with up to 10 additional septa, thick-walled,<br />

cylindrical, 2–2.5 µm wide <strong>and</strong> up to 70 µm long. Conidiogenous<br />

cells integrated, terminal, 10–30 µm long, proliferating sympodially,<br />

giving rise to a long, straight rachis with crowded, slightly darkened<br />

minute scars, about 0.5 µm diam. Conidia solitary, obovoid<br />

to fusiform with the widest part below the middle, thin-walled,<br />

verruculose, aseptate, pale brown, slightly rounded at the apex,<br />

truncate at the base, (4–)5–6(–8.5) × 2–2.5(–3) µm, with a slightly<br />

thickened <strong>and</strong> darkened hilum, 1–1.5 µm diam.<br />

Cultural characteristics: Colonies on MEA slow-growing, reaching<br />

6 mm diam after 14 d at 24 °C, velvety to hairy, colonies with entire<br />

margin, surface dark olivaceous-grey; black gelatinous exudate<br />

droplets produced on OA.<br />

Specimen examined: Brazil, São Paulo, Peruibe, Jureia Ecological Reserve, forest<br />

soil, Jan. 1991, D. Attili, holotype <strong>CBS</strong> H-19929, culture ex-type <strong>CBS</strong> 283.92.<br />

Ramichloridium cerophilum (Tubaki) de Hoog, Stud. Mycol. 15:<br />

74. 1977. Fig. 14.<br />

Basionym: Acrotheca cerophila Tubaki, J. Hattori Bot. Lab. 20: 143.<br />

1958.<br />

≡ <strong>Cladosporium</strong> cerophilum (Tubaki) Matsush., in Matsushima, Icon.<br />

Microfung. Matsush. lect. (Kobe): 34. 1975.<br />

In vitro: Submerged hyphae pale olivaceous-brown, smooth or<br />

slightly rough, 1.5–3 µm wide; aerial hyphae olivaceous-brown,<br />

smooth or slightly rough, somewhat narrower <strong>and</strong> darker than the<br />

submerged hyphae. Conidiophores unbranched, arising vertically<br />

from creeping aerial hyphae, dark brown, thick-walled, smooth or<br />

verruculose, hardly tapering towards the apex, 2–3 µm wide, up<br />

to 50 µm long, with up to 3 additional septa. Conidiogenous cells<br />

integrated, terminal, proliferating sympodially, rachis short <strong>and</strong><br />

straight, with crowded, prominent, pigmented unthickened scars,<br />

minute, approx. 0.5 µm diam. Conidia solitary, fusiform to clavate,<br />

thin-walled, smooth, 0(–1)-septate, subhyaline, (4–)6–7(–11) × (2–)<br />

2.5(–3) µm, with a conspicuous hilum, about 0.5 µm diam, slightly<br />

raised with an inconspicuous marginal frill, somehow resembling<br />

those of <strong>Cladosporium</strong>. Conidia sometimes producing 1–3(–4)<br />

short secondary conidia.<br />

Cultural characteristics: Colonies on MEA rather slow-growing,<br />

reaching 12 mm diam after 14 d at 24 °C, velvety to hairy, with<br />

entire margin; surface dark olivaceous-grey, with black gelatinous<br />

exudate droplets on OA.<br />

Specimen examined: Japan, isolated from Sasa sp., K. Tubaki, <strong>CBS</strong> 103.59, extype.<br />

72


Ramichloridium <strong>and</strong> allied genera<br />

Fig. 15. Ramichloridium indicum (<strong>CBS</strong> 171.96). A–B. Macronematous conidiophores. C–E. Sympodially proliferating conidiogenous cells, resulting in a conidium-bearing rachis<br />

with pigmented <strong>and</strong> thickened scars. F. Conidia. Scale bar = 10 µm.<br />

Fig. 16. Ramichloridium strelitziae (<strong>CBS</strong> 121711). A–C. Conidial apparatus at different stages of development, resulting in macronematous conidiophores <strong>and</strong> sympodially<br />

proliferating conidiogenous cells. D–E. Rachis with crowded, slightly pigmented, thickened, circular scars. F. Conidia. Scale bars = 10 µm.<br />

Notes: Phylogenetically, this species together with Ramichloridium<br />

apiculatum <strong>and</strong> R. musae cluster within the Mycosphaerellaceae<br />

clade. Ramichloridium cerophilum can be distinguished from its<br />

relatives by the production of secondary conidia <strong>and</strong> its distinct<br />

conidial hila.<br />

Ramichloridium indicum (Subram.) de Hoog, Stud. Mycol. 15:<br />

70. 1977. Fig. 15.<br />

Basionym: Chloridium indicum Subram., Proc. Indian Acad. Sci.,<br />

Sect. B, 42: 286. 1955 [non Rhinocladiella indica Agarwal, Lloydia<br />

32: 388. 1969].<br />

≡ Veronaea indica (Subram.) M.B. Ellis, in Ellis, More Dematiaceous<br />

Hyphomycetes: 209. 1976.<br />

= Veronaea verrucosa Geeson, Trans. Brit. Mycol. Soc. 64: 349. 1975.<br />

In vitro: Submerged hyphae smooth, thin-walled, hyaline, 1–2.5<br />

µm wide, with thin septa; aerial hyphae coarsely verrucose,<br />

olivaceous-green, rather thick-walled, 2–2.5 µm wide, with thin<br />

septa. Conidiophores arising vertically from creeping hyphae at right<br />

angles, straight, unbranched, thick-walled, smooth, dark brown,<br />

with up to 10 thin septa, up to 250 µm long, 2–4 µm wide, often<br />

with inflated basal cells. Conidiogenous cells terminally integrated,<br />

up to 165 µm long, smooth, dark brown, sympodially proliferating,<br />

rachis straight or flexuose, geniculate or nodose, subhyaline; scars<br />

thickened <strong>and</strong> darkened, clustered at nodes, approx. 0.5 µm diam.<br />

Microcyclic conidiation observed in culture. Conidia solitary, (0–)1-<br />

septate, not constricted at the septum, subhyaline to pale brown,<br />

smooth or coarsely verrucose, rather thin-walled, broadly ellipsoidal<br />

to globose, (5–)7–8(–10) × (4–)6–6.5(–9) µm, with truncate base;<br />

hilum conspicuous, slightly darkened, not thickened, about 1 µm<br />

diam.<br />

www.studiesinmycology.org<br />

73


Arzanlou et al.<br />

Cultural characteristics: Colonies on MEA reaching 35 mm diam<br />

after 14 d at 24 °C. Colonies velvety, rather compact, slightly<br />

elevated, with entire, smooth, whitish margin, dark olivaceousgreen<br />

in the central part.<br />

Specimen examined: Living culture, Feb. 1996, L. Marvanová, <strong>CBS</strong> 171.96.<br />

Ramichloridium pini de Hoog & Rahman, Trans. Brit. Mycol. Soc.<br />

81: 485. 1983.<br />

Specimen examined: U.K., Scotl<strong>and</strong>, Old Aberdeen, branch of Pinus contorta<br />

(Pinaceae), 1982, M.A. Rahman, ex-type strain, <strong>CBS</strong> 461.82 = MUCL 28942.<br />

Note: <strong>The</strong> culture examined (<strong>CBS</strong> 461.82) was sterile. For a full<br />

description see de Hoog et al. (1983).<br />

Ramichloridium strelitziae Arzanlou, W. Gams & Crous, sp. nov.<br />

MycoBank MB504551. Figs 16–17A.<br />

Etymology: Named after its host, Strelitzia.<br />

Ab aliis speciebus Ramichloridii conidiophoris brevibus, ad 40 μm longis, et<br />

cicatricibus rotundis, paulo protrudentibus distinguendum.<br />

In vitro: Submerged hyphae smooth, hyaline, thin-walled, 2–2.5 µm<br />

wide; aerial hyphae pale brown, verrucose. Conidiophores arising<br />

vertically from creeping aerial hyphae, clearly differentiated from the<br />

vegetative hyphae, subhyaline, later becoming pale brown, thickwalled,<br />

smooth or verruculose, with 1–3 additional septa; up to 40<br />

µm long <strong>and</strong> 2 µm wide. Conidiogenous cells integrated, terminal,<br />

cylindrical, variable in length, 10–35 µm long, subhyaline, later<br />

turning pale brown, fertile part as wide as the basal part, proliferating<br />

sympodially, forming a straight rachis with slightly thickened <strong>and</strong><br />

darkened, circular, somewhat protruding scars, approx. 0.5 µm<br />

diam. Conidia solitary, aseptate, smooth or verruculose, subhyaline,<br />

oblong, ellipsoidal to clavate, (3–)4–5(–5.5) × (1–)2(–2.5) µm, with<br />

truncate base <strong>and</strong> unthickened, non-pigmented hilum.<br />

Cultural characteristics: Colonies on MEA slow-growing, reaching<br />

5 mm diam after 14 d at 24 °C, with entire margin; aerial mycelium<br />

rather compact, raised, dense, olivaceous-grey; reverse olivaceousblack.<br />

Specimen examined: South Africa, KwaZulu-Natal, Durban, near Réunion,<br />

on leaves of Strelitzia nicolai, 5 Feb. 2005, W. Gams & H. Glen, <strong>CBS</strong>-H 19776,<br />

holotype, culture ex-type <strong>CBS</strong> 121711.<br />

Zasmidium Fr., Summa Veg. Sc<strong>and</strong>. 2: 407. 1849.<br />

Fig. 17. A. Ramichloridium strelitziae (<strong>CBS</strong> 121711). B. Veronaea japonica (<strong>CBS</strong><br />

776.83). C. Veronaeopsis simplex (<strong>CBS</strong> 588.66). Scale bar = 10 µm.<br />

In vitro: Submerged hyphae smooth, thin-walled, hyaline, with<br />

thin septa; aerial hyphae coarsely verrucose, olivaceous-green,<br />

thick-walled, with thin septa. Conidiophores not differentiated<br />

from vegetative hyphae, often reduced to conidiogenous<br />

Fig. 18. Zasmidium cellare (<strong>CBS</strong> 146.36). A–D. Micronematous conidiophores with terminal, integrated conidiogenous cells. E. Conidiogenous cell with pigmented, thickened<br />

<strong>and</strong> refractive scars. F–G. Primary <strong>and</strong> secondary conidia. Scale bar = 10 µm.<br />

74


Ramichloridium <strong>and</strong> allied genera<br />

Fig. 19. Rhinocladiella anceps (<strong>CBS</strong> 181.65). A. Macronematous conidiophores. B–D. Conidial apparatus at different stages of development, resulting in semi-micronematous<br />

conidiophores <strong>and</strong> sympodially proliferating conidiogenous cells. E. Conidiogenous loci. F. Conidia. Scale bars = 10 µm.<br />

cells. Conidiogenous cells integrated, predominantly terminal,<br />

sometimes lateral, arising from aerial hyphae, cylindrical, pale<br />

brown; polyblastic, proliferating sympodially producing crowded,<br />

conspicuously pigmented, almost flat, darkened, somewhat<br />

refractive scars. Conidia in short chains, cylindrical to fusiform,<br />

verrucose, obovate to obconical, pale brown, base truncate,<br />

with a conspicuous, slightly pigmented, thickened <strong>and</strong> refractive<br />

hilum. Primary conidia sometimes larger, subhyaline, verrucose<br />

or smooth-walled, 0–4-septate, variable in length, fusiform to<br />

cylindrical; conidial secession schizolytic.<br />

Type species: Zasmidium cellare (Pers. : Fr.) Fr., Summa Veg.<br />

Sc<strong>and</strong>. 2: 407. 1849.<br />

Zasmidium cellare (Pers. : Fr.) Fr., Summa Veg. Sc<strong>and</strong>. 2: 407.<br />

1849. Fig. 18.<br />

Basionym: Racodium cellare Pers., Neues Mag. Bot. 1: 123. 1794.<br />

≡ Antennaria cellaris (Pers. : Fr.) Fr., Syst. Mycol. 3: 229. 1829.<br />

≡ <strong>Cladosporium</strong> cellare (Pers. : Fr.) Sch<strong>and</strong>erl, Zentralbl. Bakteriol., 2.<br />

Abt., 94: 117. 1936.<br />

≡ Rhinocladiella cellaris (Pers. : Fr.) M.B. Ellis, in Ellis, Dematiaceous<br />

Hyphomycetes: 248. 1971.<br />

In vitro: Submerged hyphae smooth, thin-walled, hyaline, 2–3<br />

µm wide, with thin septa; aerial hyphae coarsely verrucose,<br />

olivaceous-green, rather thick-walled, 2–2.5 µm wide, with thin<br />

septa. Conidiophores not differentiated from vegetative hyphae,<br />

often reduced to conidiogenous cells. Conidiogenous cells<br />

integrated, predominantly terminal, sometimes lateral, arising from<br />

aerial hyphae, cylindrical, 20–60 µm long <strong>and</strong> 2–2.5 µm wide, pale<br />

brown, proliferating sympodially producing crowded, conspicuously<br />

pigmented scars that are thickened <strong>and</strong> refractive, about 1 µm diam.<br />

Conidia cylindrical to fusiform, verrucose, obovate to obconical,<br />

pale brown, with truncate base, (6–)9–14(–27) × 2–2.5 µm, with<br />

a conspicuous, slightly pigmented, refractive hilum, approx. 1 µm<br />

diam. Primary conidia sometimes subhyaline, verrucose or smoothwalled,<br />

thin-walled, 0–1(–4)-septate, variable in length, fusiform to<br />

cylindrical.<br />

www.studiesinmycology.org<br />

Cultural characteristics: Colonies reaching 7 mm diam after 14 d<br />

at 24 °C. Colonies velvety, rather compact, slightly elevated with<br />

entire margin; surface dark olivaceous-green in the central part,<br />

margin smooth, whitish.<br />

Specimen examined: Wall in wine cellar, Jun. 1936, H. Sch<strong>and</strong>erl, ATCC 36951<br />

= IFO 4862 = IMI 044943 = LCP 52.402 = LSHB BB274 = MUCL 10089 = <strong>CBS</strong><br />

146.36.<br />

Notes: <strong>The</strong> name Racodium Fr., typified by Ra. rupestre Pers. : Fr.,<br />

has been conserved over the older one by Persoon, with Ra. cellare<br />

as type species. De Hoog (1979) defended the use of Zasmidium in<br />

its place for the well-known wine-cellar fungus.<br />

Morphologically Zasmidium resembles Stenella Syd., <strong>and</strong><br />

both reside in the Capnodiales, though the type of Stenella, S.<br />

araguata Syd., clusters in the Teratosphaeriaceae, <strong>and</strong> the type of<br />

Zasmidium, Z. cellare, in the Mycosphaerellaceae. When accepting<br />

anamorph genera as polyphyletic within an order, preference would<br />

be given to the well-known name Stenella over the less known<br />

Zasmidium, even though the latter name is older. Further studies<br />

are required, however, to clarify if all stenella-like taxa should be<br />

accommodated in a single <strong>genus</strong>, Stenella. If this is indeed the<br />

case, a new combination for Zasmidium cellare will be proposed<br />

in Stenella, <strong>and</strong> the latter <strong>genus</strong> will have to be conserved over<br />

Zasmidium.<br />

Chaetothyriales (Herpotrichiellaceae)<br />

<strong>The</strong> four “Ramichloridium” species residing in the Chaetothyriales<br />

clade do not differ sufficiently in morphology to separate them from<br />

Rhinocladiella (type Rh. atrovirens). Because of the pale brown<br />

conidiophores, conidiogenous cells with crowded, slightly prominent<br />

scars <strong>and</strong> the occasional presence of an Exophiala J.W. Carmich.<br />

synanamorph, Rhinocladiella is a suitable <strong>genus</strong> to accommodate<br />

them. <strong>The</strong>se four species chiefly differ from Ramichloridium in the<br />

morphology of their conidial apparatus, which is clearly differentiated<br />

from the vegetative hyphae. <strong>The</strong> appropriate combinations are<br />

therefore introduced for Ramichloridium anceps, R. mackenziei, R.<br />

fasciculatum <strong>and</strong> R. basitonum.<br />

75


Arzanlou et al.<br />

Fig. 20. Rhinocladiella basitona (<strong>CBS</strong> 101460). A–B. Semi-micronematous conidiophores with verticillate branching pattern. C–D. Sympodially proliferating conidiogenous cells,<br />

giving rise to a long rachis with slightly prominent, truncate conidium-bearing denticles. E. Intercalary conidiogenous cell. F. Conidia. Scale bars = 10 µm.<br />

<strong>The</strong> <strong>genus</strong> Veronaea (type species: V. botryosa) also resides<br />

in the Chaetothyriales clade. Veronaea can be distinguished from<br />

Rhinocladiella by the absence of exophiala-type budding cells <strong>and</strong><br />

its predominantly 1-septate conidia. Furthermore, the conidiogenous<br />

loci in Veronaea are rather flat, barely prominent.<br />

Rhinocladiella Nannf., Svensk Skogsvårdsfören. Tidskr., Häfte 32:<br />

461. 1934.<br />

In vitro: Colonies dark olivaceous-brown, slow-growing, almost<br />

moist. Submerged hyphae hyaline to pale olivaceous, smooth;<br />

aerial hyphae, if present, more darkly pigmented. Exophialatype<br />

budding cells usually present in culture. Conidial apparatus<br />

usually branched, olivaceous-brown, consisting of either slightly<br />

differentiated tips of ascending hyphae or septate, markedly<br />

differentiated conidiophores. Conidiogenous cells intercalary or<br />

terminal, polyblastic, cylindrical to acicular, with a sympodially<br />

proliferating, subdenticulate rachis; scars unthickened, nonpigmented<br />

to somewhat darkened-refractive. Conidia solitary,<br />

hyaline to subhyaline, aseptate, thin-walled, smooth, subglobose,<br />

with a slightly pigmented hilum; conidial secession schizolytic.<br />

Type species: Rh. atrovirens Nannf., Svenska Skogsvårdsfören.<br />

Tidskr. 32: 461. 1934.<br />

Rhinocladiella anceps (Sacc. & Ellis) S. Hughes, Canad. J. Bot.<br />

36: 801. 1958. Fig. 19.<br />

Basionym: Sporotrichum anceps Sacc. & Ellis, Michelia 2: 576.<br />

1882.<br />

= Veronaea parvispora M.B. Ellis, in Ellis, More Dematiaceous Hyphomycetes:<br />

210. 1976.<br />

Misapplied name: Chloridium minus Corda sensu Mangenot, Rev.<br />

Mycol. (Paris) 18: 137. 1953.<br />

In vitro: Submerged hyphae subhyaline, smooth, thick-walled, 2–<br />

2.5 µm wide; aerial hyphae pale brown. Swollen germinating cells<br />

often present on MEA, giving rise to an Exophiala synanamorph.<br />

Conidiophores slightly differentiated from vegetative hyphae, arising<br />

from prostrate aerial hyphae, consisting of either unbranched or<br />

loosely branched stalks, thick-walled, golden to dark-brown, up<br />

to 350 µm tall, which may have up to 15 thin, additional septa,<br />

intercalary cells 9–14 µm long. Conidiogenous cells terminal,<br />

rarely lateral, cylindrical, occasionally intercalary, variable in length,<br />

smooth, golden to dark brown at the base, paler toward the apex,<br />

later becoming inconspicuously septate, fertile part as wide as the<br />

basal part, 15–40 × 1.5–2 µm; with crowded, slightly prominent,<br />

unpigmented, conidium-bearing denticles, about 0.5 µm diam.<br />

Conidia solitary, subhyaline, thin-walled, smooth, subglobose to<br />

ellipsoidal, 2.5–4 × 2–2.5 µm, with a less conspicuous, slightly<br />

darkened hilum, less than 0.5 µm diam.<br />

Cultural characteristics: Colonies on MEA reaching 6–12 mm diam<br />

after 14 d at 24 °C, with entire, smooth, sharp margin; mycelium<br />

powdery, becoming hairy at centre; olivaceous-green to brown,<br />

reverse dark-olivaceous.<br />

Specimens examined: Canada, Ontario, Campbellville, from soil under Thuja plicata,<br />

Apr. 1965, G. L. Barron, <strong>CBS</strong> H-7715 (isoneotype); <strong>CBS</strong> H-7716 (isoneotype); <strong>CBS</strong><br />

H-7717 (isoneotype); <strong>CBS</strong> H-7718 (isoneotype); <strong>CBS</strong> H-7719 (isoneotype), ex-type<br />

strain, <strong>CBS</strong> 181.65 = ATCC 18655 = DAOM 84422 = IMI 134453 = MUCL 8233 =<br />

OAC 10215. France, from stem of Fagus sylvatica, 1953, F. Mangenot, <strong>CBS</strong> 157.54<br />

= ATCC 15680= MUCL 1081= MUCL 7992 = MUCL 15756.<br />

Notes: Rhinocladiella anceps (conidia 2.5–4 µm long) resembles<br />

Rh. phaeophora Veerkamp & W. Gams (1983) (conidia 5.5–6 µm<br />

long), but has shorter conidia.<br />

76


Ramichloridium <strong>and</strong> allied genera<br />

Fig. 21. Rhinocladiella fasciculata (<strong>CBS</strong> 132.86). A. Conidiophores. B. Sympodially proliferating conidiogenous cells, which give rise to a long rachis with slightly prominent,<br />

unthickened scars. C. Conidia. D–E. Synanamorph consisting of conidiogenous cells with percurrent proliferation. F. Conidia. Scale bars = 10 µm.<br />

Rhinocladiella basitona (de Hoog) Arzanlou & Crous, comb. nov.<br />

MycoBank MB504552. Fig. 20.<br />

Basionym: Ramichloridium basitonum de Hoog, J. Clin. Microbiol.<br />

41: 4774. 2003.<br />

In vitro: Submerged hyphae hyaline, smooth, thin-walled, 2 µm<br />

wide; aerial hyphae rather thick-walled, pale brown. Conidiophores<br />

slightly differentiated from vegetative hyphae, profusely <strong>and</strong> mostly<br />

verticillately branched, straight or flexuose, pale-brown, 2–2.5 µm<br />

wide. Conidiogenous cells terminal, variable in length, 10–100 µm<br />

long, pale brown, straight or geniculate, proliferating sympodially,<br />

giving rise to a long, 2–2.5 µm wide rachis, with slightly prominent,<br />

truncate conidium-bearing denticles, slightly darkened. Conidia<br />

solitary, hyaline, thin-walled, smooth, pyriform to clavate, with a<br />

round apex, <strong>and</strong> slightly truncate base, (1–)3–4(–5) × 1–2 µm,<br />

hilum conspicuous, slightly darkened <strong>and</strong> thickened, less than 0.5<br />

µm diam.<br />

Cultural characteristics: Colonies on MEA reaching 19 mm diam<br />

after 14 d at 24 °C, with entire, smooth, sharp margin; mycelium<br />

rather flat <strong>and</strong> slightly elevated in the centre, pale olivaceous-grey<br />

to olivaceous-grey; reverse olivaceous-black.<br />

Specimen examined: Japan, Hamamatsu, from subcutaneous lesion with fistula on<br />

knee of 70-year-old male, Y. Suzuki, ex-type culture <strong>CBS</strong> 101460 = IFM 47593.<br />

Rhinocladiella fasciculata (V. Rao & de Hoog) Arzanlou & Crous,<br />

comb. nov. MycoBank MB504553. Fig. 21.<br />

Basionym: Ramichloridium fasciculatum V. Rao & de Hoog, Stud.<br />

Mycol. 28: 39. 1986.<br />

www.studiesinmycology.org<br />

In vitro: Submerged hyphae subhyaline, smooth, thick-walled,<br />

2–2.5 µm wide; aerial hyphae pale brown. Conidiophores arising<br />

vertically from ascending hyphae in loose fascicles, unbranched or<br />

loosely branched at acute angles, cylindrical, smooth, brown <strong>and</strong><br />

thick-walled at the base, up to 220 µm long <strong>and</strong> 2–3 µm wide, with<br />

0–5 thin additional septa. Conidiogenous cells terminal, cylindrical,<br />

30–100 µm long, thin-walled, smooth, pale brown, fertile part as<br />

wide as the basal part, up to 2 µm wide, proliferating sympodially,<br />

giving rise to a rachis with hardly prominent, slightly pigmented,<br />

not thickened scars, less than 0.5 µm diam. Conidia solitary,<br />

smooth, thin-walled, subhyaline, ellipsoidal, (2.5–)4–5(–6) × 2–3<br />

µm, with truncate, slightly pigmented hilum, about 0.5 µm diam.<br />

Synanamorph forming on torulose hyphae originating from giant<br />

cells; compact heads of densely branched hyphae forming thinwalled,<br />

lateral, subglobose cells, on which conidiogenous cells are<br />

formed; conidiogenous cells proliferating percurrently, giving rise to<br />

tubular annellated zones with inconspicuous annellations, up to 12<br />

µm long, 1–1.5 µm wide. Conidia smooth, thin-walled, aseptate,<br />

subhyaline, globose, 2–2.5 µm diam.<br />

Cultural characteristics: Colonies on MEA reaching 8 mm diam after<br />

14 d at 24 °C, with entire, smooth, sharp margin; mycelium velvety,<br />

becoming farinose in the centre due to abundant sporulation,<br />

olivaceous-green to brown, reverse dark olivaceous. Blackish<br />

droplets often produced at the centre, which contain masses of<br />

Exophiala conidia.<br />

Specimen examined: India, Karnataka, Thirathahalli, isolated by V. Rao from<br />

decayed wood, holotype <strong>CBS</strong>-H 3866, culture ex-type <strong>CBS</strong> 132.86.<br />

77


Arzanlou et al.<br />

Fig. 22. Rhinocladiella mackenziei (<strong>CBS</strong> 368.92). A. Intercalary conidiogenous cell. B–E. Semi-micronematous conidiophores <strong>and</strong> sympodially proliferating conidiogenous cells,<br />

resulting in a rachis with slightly prominent, unthickened scars. F. Conidia. Scale bar = 10 µm.<br />

78


Ramichloridium <strong>and</strong> allied genera<br />

Fig. 24. Thysanorea papuana (<strong>CBS</strong> 212.96), periconiella-like synanamorph. A. Macronematous conidiophores. B–C. Conidiophores with dense apical branches. D. Branches<br />

with different levels of branchlets. E–I. Conidiogenous cells at different stages of development; sympodially proliferating conidiogenous cells give rise to a denticulate rachis.<br />

J–K. Conidia. Scale bars = 10 µm.<br />

Fig. 23. (Page 78). Thysanorea papuana (<strong>CBS</strong> 212.96). A. Intercalary conidiogenous cell. B–I. Semi-micronematous conidiophores <strong>and</strong> sympodially proliferating conidiogenous<br />

cells, resulting in a rachis with prominent conidium bearing denticles. J–K. Microcyclic conidiation observed in slide cultures. L. Conidia. Scale bar = 10 µm.<br />

www.studiesinmycology.org<br />

79


Arzanlou et al.<br />

Rhinocladiella mackenziei (C.K. Campb. & Al-Hedaithy) Arzanlou<br />

& Crous, comb. nov. MycoBank MB504554. Fig. 22.<br />

Basionym: Ramichloridium mackenziei C.K. Campb. & Al-Hedaithy,<br />

J. Med. Veterin. Mycol. 31: 330. 1993.<br />

In vitro: Submerged hyphae subhyaline, smooth, thin-walled, 2–3<br />

µm wide; aerial hyphae pale brown, slightly narrower. Conidiophores<br />

slightly or not differentiated from vegetative hyphae, arising<br />

laterally from aerial hyphae, with one or two additional septa, often<br />

reduced to a discrete or intercalary conidiogenous cell, pale-brown,<br />

10–25 × 2.5–3.5 µm. Conidiogenous cells terminal or intercalary,<br />

variable in length, 5–15 µm long <strong>and</strong> 3–5 µm wide, occasionally<br />

slightly wider than the basal part, pale brown, rachis with slightly<br />

prominent, unpigmented, non-thickened scars, about 0.5 µm diam.<br />

Conidia golden-brown, thin-walled, smooth, ellipsoidal to obovate,<br />

subcylindical, (5–)8–9(–12) × (2–)3–3.5(–5) µm, with darkened,<br />

inconspicously thickened, protuberant or truncate hilum, less than<br />

1 µm diam.<br />

Cultural characteristics: Colonies on MEA reaching 5 mm diam<br />

after 14 d at 24 °C, with entire, smooth, sharp margin; mycelium<br />

densely lanose <strong>and</strong> elevated in the centre, olivaceous-green to<br />

brown; reverse dark olivaceous.<br />

Specimens examined: Israel, Haifa, isolated from brain abscess, <strong>CBS</strong> 368.92 =<br />

UTMB 3170; human brain abscess, E. Lefler, <strong>CBS</strong> 367.92 = NCPF 2738 = UTMB<br />

3169. Saudi Arabia, from phaeohyphomycosis of the brain, S.S.A. Al-Hedaithy,<br />

ex-type strain, <strong>CBS</strong> 650.93 = MUCL 40057 = NCPF 2808; from brain abscess,<br />

Pakistani male who travelled to Saudi Arabia, <strong>CBS</strong> 102592 = NCPF 7460. United<br />

Arab Emirates, from fatal brain abscess, <strong>CBS</strong> 102590 = NCPF 2853.<br />

Notes: Morphologically Rhinocladiella mackenziei is somewhat<br />

<strong>similar</strong> to Pleurothecium obovoideum (Matsush.) Arzanlou &<br />

Crous, which was originally isolated from dead wood. However, P.<br />

obovoideum has distinct conidiophores, <strong>and</strong> the ascending hyphae<br />

are thick-walled, <strong>and</strong> the denticles cylindrical, up to 1.5 µm long.<br />

In contrast, Rh. mackenziei has only slightly prominent denticles.<br />

Rhinocladiella mackenziei is a member of the Chaetothyriales,<br />

while P. obovoideum clusters in the Chaetosphaeriales.<br />

Thysanorea Arzanlou, W. Gams & Crous, gen. nov. MycoBank<br />

MB504555.<br />

Etymology: (Greek) thysano = brush, referring to the brush-like<br />

branching pattern, suffix derived from Veronaea.<br />

Veronaeae similis sed conidiophoris partim Periconiae similibus dense ramosis<br />

distinguenda.<br />

In vitro: Submerged hyphae subhyaline, smooth, thin-walled; aerial<br />

hyphae pale brown, smooth or verrucose. Conidiophores dimorphic;<br />

micronematous conidiophores slightly differentiated from vegetative<br />

hyphae, branched or simple, multiseptate. Conidiogenous cells<br />

terminal, polyblastic, variable in length, smooth, golden- to dark<br />

brown at the base, paler towards the apex, later sometimes<br />

inconspicuously septate; fertile part often wider than the basal<br />

part, clavate to doliiform, with crowded, more or less prominent<br />

conidium-bearing denticles, unpigmented, but slightly thickened.<br />

Macronematous conidiophores consisting of well-differentiated,<br />

thick-walled, dark brown stalks; apically repeatedly densely<br />

branched, forming a complex head, each branchlet giving rise<br />

to a conidium-bearing denticulate rachis with slightly pigmented,<br />

thickened scars. Conidia of both kinds of conidiophore formed<br />

singly, smooth, pale brown, obovoidal to pyriform, (0–)1-septate,<br />

with a truncate base <strong>and</strong> darkened hilum; conidial secession<br />

schizolytic.<br />

Type species: Thysanorea papuana (Aptroot) Arzanlou, W. Gams<br />

& Crous, comb. nov.<br />

Thysanorea papuana (Aptroot) Arzanlou, W. Gams & Crous,<br />

comb. nov. MycoBank MB504556. Figs 7C, 23–24.<br />

Basionym: Periconiella papuana Aptroot, Nova Hedwigia 67: 491.<br />

1998.<br />

In vitro: Submerged hyphae subhyaline, smooth, thin-walled, 1.5–3<br />

µm wide; aerial hyphae pale brown, smooth to verrucose, 1.5–2<br />

µm wide. Conidiophores dimorphic; micronematous conidiophores<br />

slightly differentiated from vegetative hyphae, branched or simple,<br />

up to 6-septate. Conidiogenous cells terminal or intercalary, variable<br />

in length, 5–20 µm long, thin-walled, smooth, golden- to dark brown<br />

at the base, paler toward the apex, later sometimes becoming<br />

inconspicuously septate, fertile part wider than basal part, often<br />

clavate, with crowded, more or less prominent conidium-bearing<br />

denticles, about 1 µm diam, unpigmented but slightly thickened.<br />

Conidia solitary, subhyaline, thin-walled, smooth, cylindrical to<br />

pyriform, rounded at the apex <strong>and</strong> truncate at the base, pale brown,<br />

(0–)1-septate, (5–)7–8(–11) × (2–)3(–4) µm, with a truncate base<br />

<strong>and</strong> darkened hilum, 1 µm diam. Macronematous conidiophores<br />

present in old cultures after 1 mo of incubation, consisting of welldifferentiated,<br />

thick-walled, dark brown stalks, up to 220 µm long,<br />

(4–)5–6(–7) µm wide, with up to 15 additional septa, often with<br />

inflated basal cells; apically densely branched, forming a complex<br />

head, with up to five levels of branchlets, 20–50 µm long, each<br />

branchlet giving rise to a denticulate conidium-bearing rachis; scars<br />

slightly pigmented, thickened, about 1 µm diam. Conidia solitary,<br />

thin-walled, smooth, pale brown, obovoidal to pyriform, (0–)1-<br />

septate, (4–)5–6(–8) × (2–)3(–4) µm, with a truncate base <strong>and</strong><br />

darkened hilum, 1–2 µm diam.<br />

Cultural characteristics: Colonies on MEA reaching 10 mm diam<br />

after 14 d at 24 °C, with entire, sharp margin; mycelium velvety,<br />

elevated, with colonies up to 2 mm high, surface olivaceous-grey to<br />

iron-grey; reverse greenish black.<br />

Specimen examined: Papua New Guinea, Madang Province, foothill of Finisterre<br />

range, 40.8 km along road Madang-Lae, alt. 200 m, isolated from unknown stipe, 2<br />

Nov. 1995, A. Aptroot, holotype <strong>CBS</strong>-H 6351, culture ex-type <strong>CBS</strong> 212.96.<br />

Veronaea Cif. & Montemart., Atti Ist. Bot. Lab. Crittog. Univ. Pavia,<br />

sér. 5, 15: 68. 1957.<br />

In vitro: Colonies velvety, pale olivaceous-brown, moderately<br />

fast-growing. Submerged hyphae hyaline to pale olivaceous,<br />

smooth; aerial hyphae, more darkly pigmented. Exophiala-type<br />

budding cells absent in culture. Conidiophores erect, straight or<br />

flexuose, unbranched or occasionally loosely branched, sometimes<br />

geniculate, smooth-walled, pale to medium- or olivaceous-brown.<br />

Conidiogenous cells terminally integrated, polyblastic, occasionally<br />

intercalary, cylindrical, pale brown, later often becoming septate,<br />

fertile part subhyaline, often as wide as the basal part, rachis with<br />

crowded, flat to slightly prominent, faintly pigmented, unthickened<br />

scars. Conidia solitary, smooth, cylindrical to pyriform, rounded<br />

at the apex <strong>and</strong> truncate at the base, pale brown, 1(–2)-septate;<br />

conidial secession schizolytic.<br />

Type species: Veronaea botryosa Cif. & Montemart., Atti Ist. Bot.<br />

Lab. Crittog. Univ. Pavia, sér. 5, 15: 68. 1957.<br />

Veronaea botryosa Cif. & Montemart., Atti Ist. Bot. Lab. Crittog.<br />

Univ. Pavia, sér. 5, 15: 68. 1957. Fig. 25.<br />

80


Ramichloridium <strong>and</strong> allied genera<br />

Fig. 25. Veronaea botryosa (<strong>CBS</strong> 254.57). A–C. Semi-micronematous conidiophores <strong>and</strong> sympodially proliferating conidiogenous cells. D–E. Rachis with crowded <strong>and</strong> flat<br />

scars. F–G. Microcyclic conidiation. H. Conidia. Scale bars = 10 µm.<br />

In vitro: Submerged hyphae hyaline to pale olivaceous, smooth;<br />

aerial hyphae more darkly pigmented. Conidiophores erect,<br />

straight or flexuose, unbranched or occasionally loosely branched,<br />

sometimes geniculate, smooth-walled, pale brown to olivaceousbrown,<br />

2–3 µm wide <strong>and</strong> up to 200 µm long. Conidiogenous cells<br />

terminal, occasionally intercalary, cylindrical, 10–100 µm long,<br />

pale brown, later often becoming septate, fertile part subhyaline,<br />

often as wide as the basal part, rachis with crowded, flat to slightly<br />

prominent, faintly pigmented, unthickened scars. Conidia solitary,<br />

smooth, cylindrical to pyriform, (3–)6.5–8.5(–12) × (1.5–)2–2.5(–3)<br />

µm, rounded at the apex <strong>and</strong> truncate at the base, pale brown,<br />

1(–2)-septate, with a faintly darkened, unthickened hilum, about 0.5<br />

µm diam.<br />

Cultural characteristics: Colonies on MEA reaching 30 mm diam<br />

after 14 d at 24 °C, with entire, sharp margin; mycelium velvety,<br />

slightly elevated in the centre, surface olivaceous-grey to greyishbrown;<br />

reverse greenish black.<br />

Specimens examined: India, Ramgarh, about 38 km from Jaipur, isolated from goat<br />

dung, 1 Sep. 1963, B.C. Lodha, <strong>CBS</strong> 350.65 = IMI 115127 = MUCL 7972. Italy,<br />

Tuscany, Pisa, isolated from Sansa olive slag, 1954, O. Verona, ex-type strain, <strong>CBS</strong><br />

254.57 = IMI 070233 = MUCL 9821.<br />

Veronaea compacta Papendorf, Bothalia 12: 119. 1976. Fig. 26.<br />

In vitro: Submerged hyphae subhyaline, smooth, thin-walled,<br />

1.5–3 µm wide; aerial hyphae rather thick-walled, pale brown.<br />

Conidiophores slightly differentiated from vegetative hyphae,<br />

lateral or occasionally terminal, often wider than the supporting<br />

hypha, up to 4 µm wide, unbranched or branched at acute angles,<br />

with 1–3 adititional septa, cells often inflated <strong>and</strong> flask-shaped,<br />

pale-brown, up to 60 µm long. Conidiogenous cells terminal,<br />

occasionally intercalary, variable in length, up to 10 µm long,<br />

pale brown, cylindrical to doliiform or flask-shaped, with hardly<br />

prominent denticles; scars flat, slightly pigmented, not thickened,<br />

about 0.5 µm diam. Conidia solitary, pale brown, smooth, thinwalled,<br />

ellipsoidal to ovoid, 0–1(–2)-septate, often constricted at<br />

the septa, (4–)6–7(–9) × 2–3 µm, with a round apex <strong>and</strong> truncate<br />

base; hilum prominent, slightly darkened, unthickened, about 0.5<br />

µm diam.<br />

Cultural characteristics: Colonies rather slow growing, reaching 15<br />

mm diam on MEA after 14 d at 24 °C; surface velvety to lanose,<br />

slightly raised in the centre, pale grey to pale brownish grey; reverse<br />

dark grey.<br />

Specimen examined: South Africa, soil, M.C. Papendorf, ex-type culture <strong>CBS</strong><br />

268.75.<br />

www.studiesinmycology.org<br />

81


Arzanlou et al.<br />

Fig. 26. Veronaea compacta (<strong>CBS</strong> 268.75). A–B. Semi-micronematous conidiophores <strong>and</strong> sympodially proliferating conidiogenous cells. C–D. Rachis with hardly prominent<br />

denticles. E. Conidia. Scale bar = 10 µm.<br />

Fig. 27. Veronaea japonica (<strong>CBS</strong> 776.83). A. Intercalary conidiogenous cells. B–D. Semi-micronematous conidiophores <strong>and</strong> sympodially proliferating conidiogenous cells. E.<br />

Conidia. F. Thick-walled, dark brown hyphal cells. Scale bar = 10 µm.<br />

Veronaea japonica Arzanlou, W. Gams & Crous, sp. nov.<br />

MycoBank MB504557. Figs 17B, 27.<br />

Etymology: Named after the country of origin, Japan.<br />

Veronaeae compactae similis, sed cellulis inflatis, aggregatis, crassitunicatis, fuscis<br />

in vitro formatis distinguenda.<br />

In vitro: Submerged hyphae subhyaline, smooth, thin-walled, 1.5–3<br />

µm wide; aerial hyphae slightly narrower, pale brown; hyphal cells<br />

later becoming swollen, thick-walled, dark brown, often aggregated.<br />

Conidiophores slightly differentiated from aerial vegetative hyphae,<br />

lateral, or terminal, often wider than the supporting hypha, 2–3 µm<br />

wide, up to 65 µm long, unbranched or occasionally branched,<br />

82


Ramichloridium <strong>and</strong> allied genera<br />

Fig. 28. Pleurothecium obovoideum (<strong>CBS</strong> 209.95). A–C. Conidial apparatus<br />

consisting of conidiophores with sympodially proliferating conidiogenous cells as<br />

seen in slide cultures of ca. 14 d. D. Short chain of conidia. E–G. Sympodially<br />

proliferating conidiogenous cells, resulting in a short rachis with subcylindrical to<br />

cylindrical denticles. H. Conidia. Scale bar = 10 µm.<br />

pale brown, thin-walled, smooth, with 1–3 additional septa.<br />

Conidiogenous cells terminal, occasionally intercalary, variable in<br />

length, up to 15 µm long, pale brown, cylindrical to clavate, with<br />

hardly prominent denticles; scars flat, slightly pigmented, not<br />

thickened, about 0.5 µm diam. Conidia solitary, pale brown, smooth,<br />

thin-walled, ellipsoidal to ovoid, (0–)1-septate, often constricted at<br />

the septum, (6–)7–8(–10) × 2–2.5(–4) µm, with a round apex <strong>and</strong><br />

truncate base; hilum unthickened but slightly darkened, about 1 µm<br />

diam.<br />

Cultural characteristics: Colonies rather slow growing, reaching 7.5<br />

mm diam on MEA after 14 d at 24 °C; surface velvety to lanose,<br />

slightly raised in the centre, olivaceous-brown, with entire margin;<br />

reverse dark-olivaceous.<br />

Specimen examined: Japan, Kyoto, Daitokuji Temple, Kyoto, inside dead bamboo<br />

culm, Dec. 1983, W. Gams, holotype <strong>CBS</strong>-H 3490, ex-type culture <strong>CBS</strong> 776.83.<br />

Note: This species is morphologically <strong>similar</strong> to V. compacta<br />

(Papendorf 1976), but can be distinguished based on the presence<br />

of dark brown, swollen hyphal cells in culture, which are absent in<br />

V. compacta.<br />

Pleurothecium obovoideum clade (Chaetosphaeriales)<br />

Ramichloridium obovoideum was regarded as <strong>similar</strong> to<br />

“Ramichloridium” (Rhinocladiella) mackenziei by some authors,<br />

<strong>and</strong> subsequently reduced to synonymy (Ur-Rahman et al. 1988).<br />

However, R. obovoideum clusters with Carpoligna pleurothecii,<br />

the teleomorph of Pleurothecium Höhn. Because it is also<br />

morphologically <strong>similar</strong> to other species of Pleurothecium, we<br />

herewith combine it into that <strong>genus</strong>.<br />

Pleurothecium obovoideum (Matsush.) Arzanlou & Crous, comb.<br />

nov. MycoBank MB504558. Fig. 28.<br />

Basionym: Rhinocladiella obovoidea Matsush., Icones Microfung.<br />

Mats. lect.: 123. 1975.<br />

≡ Ramichloridium obovoideum (Matsush.) de Hoog, Stud. Mycol. 15: 73.<br />

1977.<br />

In vitro: Submerged hyphae smooth, hyaline, thin-walled, 1–2 µm<br />

wide; aerial hyphae hyaline to subhyaline, smooth. Conidiophores<br />

arising vertically from creeping hyphae, ascending hyphae thickwalled<br />

<strong>and</strong> dark brown; conidiophores 10–35 µm long, 1–2-septate,<br />

often reduced to a conidiogenous cell, unbranched, thick-walled,<br />

smooth, tapering towards the apex, pale brown. Conidiogenous<br />

cells integrated, cylindrical to ampulliform, 5–20 µm long, pale<br />

brown, elongating sympodially, with a short rachis giving rise to<br />

denticles, 1 µm long, slightly pigmented. Conidia aseptate, solitary<br />

or in short chains of up to 3, smooth, pale brown, ellipsoidal to<br />

obovate, (9–)11–12(–14.5) × (3–)4(–5) µm, smooth, thin-walled,<br />

with a more or less rounded apex, a truncate base <strong>and</strong> a slightly<br />

darkened, unthickened hilum, 1.5 µm diam.<br />

www.studiesinmycology.org<br />

83


Arzanlou et al.<br />

Cultural characteristics: Colonies slow-growing, reaching 15 diam<br />

after 14 d at 24 °C, with entire, smooth margin; surface rather<br />

compact, mycelium mainly flat, submerged, some floccose to<br />

lanose aerial mycelium in the centre, buff; reverse honey.<br />

Specimen examined: Japan, Kobe Municipal Arboretum, T. Matsushima, from dead<br />

leaf of Pasania edulis, <strong>CBS</strong> 209.95 = MFC 12477.<br />

Incertae sedis (Sordariomycetes)<br />

Ramichloridium schulzeri clade<br />

Ramichloridium schulzeri, including its varieties, clusters<br />

near Thyridium Nitschke <strong>and</strong> the Magnaporthaceae, <strong>and</strong> is<br />

phylogenetically as well as morphologically distinct from the other<br />

genera in the Ramichloridium complex. To accommodate these<br />

taxa, a new <strong>genus</strong> is introduced below.<br />

Myrmecridium Arzanlou, W. Gams & Crous, gen. nov. MycoBank<br />

MB504559.<br />

Etymology: (Greek) myrmekia = wart, referring to the wart-like<br />

denticles on the rachis, suffix -ridium from Chloridium.<br />

Genus ab allis generibus Ramichloridii similibus rachide recta longa, subhyalina,<br />

denticulis distantibus, verruciformibus praedita distinguendum.<br />

In vitro: Colonies moderately fast-growing, flat, with mainly<br />

submerged mycelium, <strong>and</strong> entire margin, later becoming powdery<br />

to velvety, pale orange to orange. Mycelium rather compact, mainly<br />

submerged, in the centre velvety with fertile bundles of hyphae.<br />

Conidiophores arising vertically <strong>and</strong> clearly distinct from creeping<br />

hyphae, unbranched, straight or flexuose, brown, thick-walled.<br />

Conidiogenous cells terminally integrated, polyblastic, cylindrical,<br />

straight or flexuose, pale brown, sometimes secondarily septate,<br />

fertile part subhyaline, as wide as the basal part, with scattered<br />

pimple-shaped, apically pointed, unpigmented, conidium-bearing<br />

denticles. Conidia solitary, subhyaline, smooth or finely verrucose,<br />

rather thin-walled, with a wing-like gelatinous sheath, obovoidal or<br />

fusiform, tapering towards a narrowly truncate base with a slightly<br />

prominent, unpigmented hilum; conidial secession schizolytic.<br />

Type species: Myrmecridium schulzeri (Sacc.) Arzanlou, W. Gams<br />

& Crous, comb. nov.<br />

Notes: Myrmecridium schulzeri was fully described as Acrotheca<br />

acuta Grove by Hughes (1951). <strong>The</strong> author discussed several<br />

genera, none of which is suitable for the present fungus for various<br />

reasons as analysed by de Hoog (1977). Only Gomphinaria Preuss<br />

is not yet sufficiently documented. Our examination of G. amoena<br />

Preuss (B!) showed that this is an entirely different fungus, of which<br />

no fresh material is available to ascertain its position.<br />

Myrmecridium can be distinguished from other ramichloridiumlike<br />

fungi by having entirely hyaline vegetative hyphae, <strong>and</strong> widely<br />

scattered, pimple-shaped denticles on the long hyaline rachis.<br />

<strong>The</strong> conidial sheath is visible in lactic acid mounts with brightfield<br />

microscopy. <strong>The</strong> Myrmecridium clade consists of several<br />

subclusters, which are insufficiently resolved based on the ITS<br />

sequence data. However, two morphologically distinct varieties of<br />

Myrmecridium are treated here. <strong>The</strong> status of the other isolates in<br />

this clade will be dealt with in a future study incorporating more<br />

strains, <strong>and</strong> using a multi-gene phylogenetic approach.<br />

Myrmecridium schulzeri (Sacc.) Arzanlou, W. Gams & Crous,<br />

comb. nov. MycoBank MB504560. var. schulzeri Figs 7B, 29.<br />

Basionym: Psilobotrys schulzeri Sacc., Hedwigia 23: 126. 1884.<br />

≡ Chloridium schulzerii (Sacc.) Sacc., Syll. Fung. 4: 322. 1886.<br />

≡ Rhinocladiella schulzeri (Sacc.) Matsush., Icon. Microfung. Mats. lect.<br />

(Kobe): 124. 1975.<br />

≡ Ramichloridium schulzeri (Sacc.) de Hoog, Stud. Mycol. 15: 64. 1977<br />

var. schulzeri.<br />

= Acrotheca acuta Grove, J. Bot., Lond. 54: 222. 1916.<br />

≡ Pleurophragmium acutum (Grove) M.B. Ellis in Ellis, More Dematiaceous<br />

Hyphomycetes: 165. 1976.<br />

= Rhinotrichum multisporum Doguet, Rev. Mycol., Suppl. Colon. 17: 78. 1953<br />

(nom. inval. Art. 36) [non Acrotheca multispora (Preuss) Sacc., Syll. Fung. 4:<br />

277. 1886].<br />

[non Acrothecium (?) multisporum G. Arnaud, Bull. Trimestriel Soc.<br />

Mycol. France 69: 288. 1953 (nom. inval. Art. 36)].<br />

[non Acrothecium multisporum G. Arnaud sensu Tubaki, J. Hattori<br />

Bot. Lab. 20: 145. 1958].<br />

In vitro: Submerged hyphae hyaline, thin-walled, 1–2 µm wide;<br />

aerial hyphae, if present, pale olivaceous-brown. Conidiophores<br />

arising vertically from creeping aerial hyphae, unbranched, straight,<br />

reddish brown, thick-walled, septate, up to 250 µm tall, 2.5–3.5<br />

µm wide, with 2–7 additional septa, basal cell often inflated, 3.5–5<br />

µm wide. Conidiogenous cells integrated, cylindrical, variable in<br />

length, 15–110 µm long, subhyaline to pale brown, later becoming<br />

inconspicuously septate, fertile part subhyaline, as wide as the<br />

basal part, forming a straight rachis with scattered, pimple-shaped<br />

denticles less than 1 µm long <strong>and</strong> approx. 0.5 µm wide, apically<br />

pointed, unpigmented, slightly thickened scars. Conidia solitary,<br />

subhyaline, thin-walled, smooth or finely verrucose, surrounded<br />

by a wing-like, gelatinous conidial sheath, up to 0.5 µm thick,<br />

ellipsoid, obovoid or fusiform, (6–)9–10(–12) × 3–4 µm, tapering to<br />

a subtruncate base; hilum unpigmented, inconspicuous.<br />

Cultural characteristics: Colonies reaching 29 mm diam after 14<br />

d at 24 °C, pale orange to orange, with entire margin; mycelium<br />

flat, rather compact, later becoming farinose or powdery due to<br />

sporulation, which occurs in concentric zones when incubated on<br />

the laboratory bench.<br />

Specimens examined: Germany, Kiel-Kitzeberg, from wheat-field soil, W. Gams,<br />

<strong>CBS</strong> 134.68 = ATCC 16310. <strong>The</strong> Netherl<strong>and</strong>s, isolated from a man, bronchial<br />

secretion, A. Visser, <strong>CBS</strong> 156.63 = MUCL 1079; Lienden, isolated from Triticum<br />

aestivum root, C.L. de Graaff, <strong>CBS</strong> 325.74 = JCM 7234.<br />

Myrmecridium schulzeri var. tritici (M.B. Ellis) Arzanlou, W.<br />

Gams & Crous, comb. nov. MycoBank MB504562.<br />

Basionym: Pleurophragmium tritici M.B. Ellis, in Ellis, More<br />

Dematiaceous Hyphomycetes: 165. 1976.<br />

≡ Ramichloridium schulzeri var. tritici (M.B. Ellis) de Hoog, Stud. Mycol.<br />

15: 68. 1977.<br />

Specimen examined: Irel<strong>and</strong>, Dublin, on wheat stem, Oct. 1960, J.J. Brady,<br />

holotype IMI 83291.<br />

Notes: No reliable living culture is available of this variety. Based<br />

on a re-examination of the type specimen in this study, the variety<br />

appears sufficiently distinct from Myrmecridium schulzeri var.<br />

schulzeri based on the frequent production of septate conidia.<br />

Myrmecridium flexuosum (de Hoog) Arzanlou, W. Gams & Crous,<br />

comb. et stat. nov. MycoBank MB504563. Fig. 30.<br />

Basionym: Ramichloridium schulzeri var. flexuosum de Hoog, Stud.<br />

Mycol. 15: 67. 1977.<br />

In vitro: Submerged hyphae hyaline, thin-walled, 1–2 µm wide.<br />

Conidiophores unbranched, flexuose, arising from creeping aerial<br />

84


Ramichloridium <strong>and</strong> allied genera<br />

Fig. 29. Myrmecridium schulzeri (<strong>CBS</strong> 325.74). A. Macronematous conidiophores. B. Inflated basal cells visible in some conidiophores. C–E. Conidial apparatus at different<br />

stages of development, resulting in macronematous conidiophores <strong>and</strong> sympodially proliferating conidiogenous cells. F–G. Rachis with scattered, pimple-shaped denticles.<br />

H. Conidia. Scale bars: A =100 µm, B–H = 10 µm.<br />

hyphae, pale brown, up to 250 µm tall, 3–3.5 µm wide, thick-walled,<br />

smooth, with up to 24 thin septa, delimiting 8–12 µm long cells.<br />

Conidiogenous cells integrated, elongating sympodially, cylindrical,<br />

20–150 µm long, flexuose, brown at the base, subhyaline in the<br />

upper part, later becoming inconspicuously septate; rachis slightly<br />

flexuose, subhyaline, as wide as the basal part, thick-walled near<br />

the base, hyaline <strong>and</strong> thin-walled in the apical part, with scattered<br />

pimple-shaped, unpigmented, approx. 0.5 µm long denticles.<br />

Conidia solitary, subhyaline, thin-walled, finely verrucose, with<br />

a wing-like gelatinous sheath, approx. 0.5 µm wide, ellipsoid<br />

to obovoid, (5–)6–7(–9) × 3–4 µm; hilum slightly prominent,<br />

unpigmented, approx. 0.5 µm diam.<br />

Cultural characteristics: Colonies reaching 40 mm diam after 14 d<br />

at 24 °C; mycelium submerged, flat, smooth; centrally orange, later<br />

becoming powdery to velvety <strong>and</strong> greyish brown due to sporulation,<br />

with sharp, smooth, entire margin; reverse yellowish orange.<br />

Specimen examined: Surinam, isolated from soil, J.H. van Emden, ex-type culture<br />

<strong>CBS</strong> 398.76 = JCM 6968.<br />

www.studiesinmycology.org<br />

Note: This former variety is sufficiently distinguished from M.<br />

schulzeri s. str. by its flexuose conidiophores <strong>and</strong> conidia which<br />

lack an acuminate base, to be regarded as a separate species.<br />

Ramichloridium torvi (Ellis & Everh.) de Hoog, Stud. Mycol. 15:<br />

79. 1977.<br />

≡ Ramularia torvi Ellis & Everh., Rep. Missouri Bot. Gard. 9: 119. 1898.<br />

≡ Hansfordia torvi (Ellis & Everh.) Deighton & Piroz., Mycol. Pap. 101:<br />

39. 1965.<br />

= Acladium biophilum Cif., Sydowia 10: 164. 1956.<br />

≡ Hansfordia biophila (Cif.) M.B. Ellis, in Ellis, More Dematiaceous<br />

Hyphomycetes: 199. 1976.<br />

Specimen: Jamaica, Port Marant, Dec. 1890, on leaves of Solanum torvum,<br />

holotype of Ramularia torvi (NY) (specimen not examined).<br />

Notes: According to the description <strong>and</strong> illustration of R. torvi<br />

provided by de Hoog (1977), this appears to be an additional<br />

species of Myrmecridium. Although it is morphologically <strong>similar</strong> to<br />

M. flexuosum in having a flexuose rachis, it differs from the other<br />

species of the <strong>genus</strong> by having smooth, clavate conidia. Fresh<br />

collections <strong>and</strong> cultures would be required to resolve its status.<br />

85


Arzanlou et al.<br />

Fig. 30. Myrmecridium flexuosum (<strong>CBS</strong> 398.76). A–C. Conidial apparatus at different stages of development, resulting in macronematous conidiophores with sympodially<br />

proliferating conidiogenous cells. D–H. Sympodially proliferating conidiogenous cells giving rise to a flexuose conidium-bearing rachis with pimple-shaped denticles. I. Conidia.<br />

Scale bar = 10 µm.<br />

Fig. 31. Pseudovirgaria hyperparasitica (<strong>CBS</strong> 121739). A–D. Conidial apparatus at different stages of development; conidiogenous cells with geniculate proliferation. E. Conidia.<br />

Scale bar = 10 µm.<br />

86


Ramichloridium <strong>and</strong> allied genera<br />

Pseudovirgaria H.D. Shin, U. Braun, Arzanlou & Crous, gen. nov.<br />

MycoBank MB504564.<br />

Etymology: Named after its morphological <strong>similar</strong>ity to Virgaria.<br />

Hyphomycetes. Uredinicola. Coloniae in vivo pallide vel modice brunneae,<br />

ferrugineae vel cinnamomeae, in vitro lentissime crescentes, murinae. Mycelium<br />

immersum et praecipue externum, ex hyphis ramosis et cellulis conidiogenis<br />

integratis compositum, conidiophoris ab hyphis vegetativis vix distinguendis.<br />

Hyphae ramosae, septatae, leves, tenuitunicatae, hyalinae vel pallide brunneae.<br />

Cellulae conidiogenae integratae in hyphis repentibus, terminales et intercalares,<br />

polyblasticae, sympodialiter proliferentes, subcylindricae vel geniculatae, cicatricibus<br />

conspicuis, solitariis vel numerosis, dispersis vel aggregatis, subdenticulatis,<br />

prominentibus, umbonatis vel apicem versus paulo attenuatis, non inspissatis,<br />

non vel parce fuscatis-refringentibus. Conidia solitaria, holoblastica, plus minusve<br />

obovoidea, recta vel leniter curvata, asymmetrica, continua, hyalina, subhyalina<br />

vel pallidissime olivaceo-brunnea, hilo subconspicuo vel conspicuo, truncato<br />

vel rotundato, non inspissato, non vel lenissime fuscato-refringente; secessio<br />

schizolytica.<br />

Hyperparasitic on uredosori of rust fungi. Colonies in vivo pale<br />

to medium brown, rusty or cinnamom, in vitro slow-growing,<br />

pale to dark mouse-grey. Mycelium immersed <strong>and</strong> mainly aerial,<br />

composed of branched hyphae with integrated conidiogenous<br />

cells, differentiation between vegetative hyphae <strong>and</strong> conidiophores<br />

barely possible. Hyphae branched, septate, smooth, thin-walled,<br />

hyaline to pale brown. Conidiogenous cells <strong>similar</strong>ly hyaline to<br />

pale brown, integrated in creeping threads (hyphae), terminal <strong>and</strong><br />

intercalary, polyblastic, proliferation sympodial, rachis subcylindrical<br />

to geniculate, conidiogenous loci (scars) conspicuous, solitary to<br />

numerous, scattered to aggregated, subdenticulate, bulging out,<br />

umbonate or slightly attenuated towards a rounded apex, wall<br />

unthickened, not to slightly darkened-refractive. Conidia solitary,<br />

formation holoblastic, more or less obovoid, straight to somewhat<br />

curved, asymmetrical, aseptate, hyaline, subhyaline to very pale<br />

olivaceous-brown, with more or less conspicuous hilum, truncate to<br />

rounded, unthickened, not or slightly darkened-refractive; conidial<br />

secession schizolytic.<br />

Type species: Pseudovirgaria hyperparasitica H.D. Shin, U. Braun,<br />

Arzanlou & Crous, sp.nov.<br />

Notes: Other ramichloridium-like isolates from various rust species<br />

form another unique clade, sister to Radulidium subulatum (de<br />

Hoog) Arzanlou, W. Gams & Crous <strong>and</strong> Ra. epichloës (Ellis &<br />

Dearn.) Arzanlou, W. Gams & Crous in the Sordariomycetidae.<br />

Although Pseudovirgaria is morphologically <strong>similar</strong> to Virgaria Nees,<br />

it has hyaline to pale brown hyphae, conidia <strong>and</strong> conidiogenous<br />

cells. <strong>The</strong> conidiogenous cells are integrated in creeping threads<br />

(hyphae), terminal <strong>and</strong> intercalary, <strong>and</strong> the proliferation is distinctly<br />

sympodial. <strong>The</strong> subdenticulate conidiogenous loci are scattered,<br />

solitary, at small shoulders of geniculate conidiogenous cells,<br />

caused by sympodial proliferation, or aggregated, forming slight<br />

swellings of the rachis, i.e., a typical raduliform rachis as in Virgaria<br />

is lacking. Furthermore, the conidiogenous loci of Pseudovirgaria<br />

are bulging, convex, slightly attenuated towards the rouded apex,<br />

in contrast to more cylindrical denticles in Virgaria (Ellis 1971).<br />

<strong>The</strong> scar type of Pseudovirgaria is peculiar due to its convex,<br />

papilla-like shape <strong>and</strong> reminiscent of conidiogenous loci in plantpathogenic<br />

genera like Neoovularia U. Braun <strong>and</strong> Pseudodidymaria<br />

U. Braun (Braun 1998). <strong>The</strong> superficially <strong>similar</strong> <strong>genus</strong> Veronaea is<br />

quite distinct from Pseudovirgaria by having erect conidiophores<br />

with a typical rachis <strong>and</strong> crowded conidiogenous loci which are<br />

flat or only slightly prominent <strong>and</strong> darkened. Pseudovirgaria is<br />

characterised by its mycelium which is composed of branched<br />

hyphae with integrated, terminal <strong>and</strong> intercalary conidiogenous<br />

cells. A differentiation between branched hyphae <strong>and</strong> “branched<br />

www.studiesinmycology.org<br />

conidiophores” is difficult <strong>and</strong> barely possible. It remains unclear<br />

if the “creeping threads” <strong>and</strong> terminal branches of hyphae are<br />

to be interpreted as “creeping conidiophores”. In any case, the<br />

mycelium forms complex fertile branched hyphal structures in which<br />

individual conidiophores are barely discernable. <strong>The</strong>se structures<br />

<strong>and</strong> difficulties in discerning individual conidiophores remind one of<br />

some species of Pseudocercospora Speg. <strong>and</strong> other cercosporoid<br />

genera with abundant superficial mycelium in vivo.<br />

Pseudovirgaria hyperparasitica H.D. Shin, U. Braun, Arzanlou &<br />

Crous, sp. nov. MycoBank MB504565. Figs 6A, 31.<br />

Etymology: Named after its hyperparasitic habit on rust fungi.<br />

Hyphae 1.5–4 µm latae, tenuitunicatae, ≤ 0.5 µm crassae. Cellulae conidiogenae<br />

15–50 × 2–5 µm, tenuitunicatae (≤ 0.5 µm), cicatricibus (0.5–)1.0(–1.5) µm diam,<br />

0.5–1 µm altis. Conidia saepe obovoidea, interdum subclavata, 10–20 × 5–9 µm,<br />

apice rotundato vel paulo attenuato, basi truncata vel rotundata, hilo ca 1 µm<br />

diam.<br />

In vivo: Colonies on rust sori, thin to moderately thick, loose,<br />

cobwebby, to dense, tomentose, pale to medium brown, rusty<br />

or cinnamon. Mycelium partly immersed in the sori, but mainly<br />

superficial, composed of a system of branched hyphae with<br />

integrated conidiogenous cells (fertile threads), distinction between<br />

conidiophores <strong>and</strong> vegetative hyphae difficult <strong>and</strong> barely possible.<br />

Hyphae 1.5–4 µm wide, hyaline, subhyaline to pale yellowish,<br />

greenish or very pale olivaceous, light brownish in mass, thin-walled<br />

(≤ 0.5 µm), smooth, pluriseptate, occasionally slightly constricted<br />

at the septa. Conidiogenous cells integrated in creeping fertile<br />

threads, terminal or intercalary, 15–50 µm long, 2–5 µm wide,<br />

subcylindrical to geniculate, subhyaline to very pale brownish, wall<br />

thin, ≤ 0.5 µm, smooth, proliferation sympodial, with a single to<br />

usually several conidiogenous loci per cell, often crowded, causing<br />

slight swellings, up to 6 µm wide, subdenticulate loci, formed by the<br />

slightly bulging wall, convex, slightly narrowed towards the rounded<br />

apex, (0.5–)1.0(–1.5) µm diam <strong>and</strong> 0.5–1 µm high, wall of the loci<br />

unthickened, not or slightly darkened-refractive, in surface view<br />

visible as minute circle (only rim visible <strong>and</strong> dark). Conidia solitary,<br />

obovoid, often slightly curved with ± unequal sides, 10–20 × 5–9<br />

µm, aseptate, subhyaline, pale yellowish greenish to very pale<br />

olivaceous, wall ≤ 0.5 µm thick, smooth, apex slightly attenuated<br />

to usually broadly rounded, base rounded to somewhat attenuated<br />

towards a more or less conspicuous hilum, (0.5–)1(–1.5) µm<br />

diam, convex to truncate, unthickened, not to slightly darkenedrefractive.<br />

In vitro: Submerged hyphae hyaline to subhyaline, smooth; aerial<br />

hyphae smooth, subhyaline, up to 4 µm wide. Conidiogenous cells<br />

arising imperceptably from aerial vegetative hyphae, terminal,<br />

occasionally intercalary, holoblastic, proliferating sympodially in a<br />

geniculate pattern, with more or less long intervals between groups<br />

of scars; loci slightly darkened, unthickened, approx. 0.5 µm diam.<br />

Conidia hyaline to subhyaline, aseptate, ovoid, often somewhat<br />

curved, (10–)13–15(–17) × (5–)6–7(–8) µm, with truncate base<br />

<strong>and</strong> acutely rounded apex; hila unthickened, slightly darkenedrefractive.<br />

Cultural characteristics: Colonies on MEA rather slow-growing,<br />

reaching 11 mm diam after 14 d at 24 °C, pale to dark mouse-grey,<br />

velvety, compacted, with colonies being up to 1 mm high.<br />

Specimens examined: Korea, Seoul, on uredosori of Frommeëlla sp., on Duchesnea<br />

chrysantha, 17 Sep. 2003, H.D. Shin, paratype, 4/10, CPC 10702–10703 = <strong>CBS</strong><br />

121735–121736, HAL 2053 F; Chunchon, on Phragmidium griseum on Rubus<br />

crataegifolius, 20 Jul. 2004, H.D. Shin, paratype, 2/8, HAL 2057 F; Suwon,<br />

87


Arzanlou et al.<br />

Fig. 32. Radulidium subulatum (<strong>CBS</strong> 405.76). A–B. Macronematous conidiophores with sympodially proliferating conidiogenous cells, resulting in a conidium-bearing rachis.<br />

C–D. Rachis with crowded, blunt conidium-bearing denticles. E. Conidia. Scale bar = 10 µm.<br />

Fig. 33. Radulidium epichloës (<strong>CBS</strong> 361.63). A–C. Conidial apparatus at different stages of development, resulting in semi-micronematous conidiophores <strong>and</strong> sympodially<br />

proliferating conidiogenous cells. D. Rachis with crowded, blunt conidium-bearing denticles. E–F. Conidiophores with acute branches in the lower part. G. Conidia. Scale bar<br />

= 10 µm.<br />

88


Ramichloridium <strong>and</strong> allied genera<br />

on Phragmidium pauciloculare on Rubus parvifolius, 14 Oct. 2003, H.D. Shin,<br />

paratype, 23/10, HAL 2055 F; Hongchon, on Phragmidium rosae-multiflorae on<br />

Rosa multiflora, 11 Aug. 2004, H.D. Shin, paratype, 23/8, HAL 2056 F; Yangpyong,<br />

on Phragmidium sp. on Rubus coreanus, 30 Sep. 2003, H.D. Shin, paratype, 11/10-<br />

1, CPC 10704–10705 = <strong>CBS</strong> 121737–121738, HAL 2052 F, <strong>and</strong> the same locality,<br />

23 Jul. 2004, HAL 2058 F; Chunchon, on Pucciniastrum agrimoniae on Agrimonia<br />

pilosa, 7 Oct. 2002, H.D. Shin, holotype, HAL 2054 F, culture ex-type CPC 10753–<br />

10755 = <strong>CBS</strong> 121739–121741.<br />

Radulidium subulatum <strong>and</strong> Ra. epichloës clade<br />

Ramichloridium subulatum <strong>and</strong> R. epichloës form a distinct, wellsupported<br />

clade with uncertain affinity. This clade is morphologically<br />

distinct <strong>and</strong> a new <strong>genus</strong> is introduced below to accommodate it.<br />

Radulidium Arzanlou, W. Gams & Crous, gen. nov. MycoBank<br />

MB504566.<br />

Etymology: Latin radula = A flexible tongue-like organ in gastropods,<br />

referring to the radula-like denticles on the rachis.<br />

Genus ab aliis generibus Ramichloridii similibus denticulis densissimis, prominentibus,<br />

hebetibus in rachide e cellula conidiogena aculeata orta distinguendum.<br />

Type species: Radulidium subulatum (de Hoog) Arzanlou, W. Gams<br />

& Crous, comb. nov.<br />

In vitro: Colonies fast-growing, velvety, floccose near the margin,<br />

centrally with fertile hyphal bundles up to 10 mm high, about 2<br />

mm diam, with entire but vague margin; mycelium whitish, later<br />

becoming greyish brown. Submerged hyphae smooth, thin-walled.<br />

Conidiophores usually reduced to polyblastic conidiogenous cells<br />

arising from undifferentiated or slightly differentiated aerial hyphae,<br />

terminally integrated or lateral, rarely a branched conidiophore<br />

present, smooth, slightly thick-walled, pale brown, cylindrical to<br />

acicular, widest at the base <strong>and</strong> tapering towards the apex; apical<br />

part forming a pale brown, generally straight rachis, with crowded,<br />

prominent, blunt denticles, suggesting a gastropod radula; denticles<br />

0.5–1 µm long, apically pale brown. Conidia solitary, subhyaline,<br />

thin- or slightly thick-walled, smooth or verrucose, obovoidal,<br />

fusiform to subcylindrical, base subtruncate <strong>and</strong> with a slightly<br />

prominent, conspicuously pigmented hilum; conidial secession<br />

schizolytic.<br />

Notes: Radulidium can be distinguished from other ramichloridiumlike<br />

fungi by its slightly differentiated conidiophores <strong>and</strong> prominent,<br />

blunt, very dense conidium-bearing denticles. Although the<br />

Radulidium clade consists of several subclusters that correlate<br />

with differences in morphology, the ITS sequence data appear<br />

insufficient to resolve this species complex. <strong>The</strong>refore, only two<br />

species of Radulidium with clear morphological <strong>and</strong> molecular<br />

differences are treated here. <strong>The</strong> phylogenetic situation of other<br />

taxa in this clade will be treated in a further study employing a multigene<br />

approach.<br />

Radulidium subulatum (de Hoog) Arzanlou, W. Gams & Crous,<br />

comb. nov. MycoBank MB504567. Figs 10C, 32.<br />

Basionym: Ramichloridium subulatum de Hoog, Stud. Mycol. 15:<br />

83. 1977.<br />

Misapplied name: Rhinocladiella elatior Mangenot sensu dal Vesco<br />

& B. Peyronel, Allionia 14: 38. 1968.<br />

In vitro: Submerged hyphae hyaline, thin-walled, 1–2.5 µm wide;<br />

aerial hyphae brownish. Conidiogenous cells arising laterally from<br />

vegetative hyphae, pale brown, smooth, thick-walled, sometimes<br />

without a basal septum, cylindrical to aculeate, tapering gradually<br />

towards the apex, widest at the base, 25–40 × 2–3 µm; proliferating<br />

sympodially, forming a pale brown rachis, with densely crowded,<br />

prominent, blunt conidium-bearing denticles, with pale brown apex.<br />

Conidia solitary, subhyaline, thin-walled, smooth, ellipsoidal to<br />

almost clavate, 5–7 × 1.5–2 µm, with a slightly pigmented, nonrefractive<br />

hilum, about 1 µm diam.<br />

Cultural characteristics: Colonies on MEA rather fast growing,<br />

reaching 50 mm diam after 14 d at 24 °C, with entire but vague<br />

margin, velvety, floccose near the margin, centrally with fertile<br />

hyphal bundles up to 10 mm high, about 2 mm diam; mycelium<br />

whitish, later becoming greyish brown; reverse grey, zonate.<br />

Specimens examined: Czech Republic, on Phragmites australis, A. Samšiňáková,<br />

ex-type culture <strong>CBS</strong> 405.76; Opatovicky pond, from Lasioptera arundinis (gall<br />

midge) mycangia on Phragmites australis, M. Skuhravá, <strong>CBS</strong> 101010.<br />

Radulidium epichloës (Ellis & Dearn.) Arzanlou, W. Gams &<br />

Crous, comb. nov. MycoBank MB504568. Fig. 33.<br />

Basionym: Botrytis epichloës Ellis & Dearn., Canad. Record Sci.<br />

9: 272. 1893.<br />

≡ Ramichloridium epichloës (Ellis & Dearn.) de Hoog, Stud. Mycol. 15:<br />

81. 1977.<br />

In vitro: Submerged hyphae hyaline, thin-walled, 1–2.5 µm wide;<br />

aerial hyphae somewhat darker. Conidiogenous cells arising<br />

laterally or terminally from undifferentiated or slightly differentiated<br />

aerial hyphae, occasionally acutely branched in the lower part,<br />

smooth, thick-walled, pale brown, more or less cylindrical, later<br />

with thin septa, 25–47 µm long; proliferating sympodially, forming<br />

a rather short, pale brown, straight or somewhat geniculate<br />

rachis, with crowded, prominent, blunt denticles with pale brown<br />

apex. Conidia solitary, subhyaline, rather thin-walled, verruculose,<br />

obovoidal to fusiform, (4.5–)7–8(–11) × 2–3 µm, with a pigmented<br />

hilum, 1–1.5 µm diam.<br />

Cultural characteristics: Colonies reaching 45 mm diam after 14 d<br />

at 24 °C, with smooth, rather vague, entire margin; velvety, centrally<br />

floccose <strong>and</strong> elevated up to 2 mm high; surface mycelium whitish,<br />

later becoming greyish brown; reverse pale ochraceous.<br />

Specimen examined: U.S.A., Cranberry Lake, Michigan, isolated from Epichloë<br />

typhina on Glyceria striata, G.L. Hennebert, <strong>CBS</strong> 361.63 = MUCL 3124; specimen<br />

in MUCL designated here as epitype.<br />

Veronaea-like clade, allied to the Annulatascaceae<br />

A veronaea-like isolate from Bertia moriformis clusters near the<br />

Annulatascaceae, <strong>and</strong> is morphologically distinct from other known<br />

anamorph genera in the Ramichloridium complex, <strong>and</strong> therefore a<br />

new <strong>genus</strong> is introduced to accommodate it.<br />

Rhodoveronaea Arzanlou, W. Gams & Crous, gen. nov. MycoBank<br />

MB504569.<br />

Etymology: (Greek) rhodon = the rose, referring to the red-brown<br />

conidiophores, suffix -veronaea from Veronaea.<br />

Genus ab aliis generibus Ramichloridii similibus basi condiorum late truncata et<br />

marginata distinguenda.<br />

In vitro: Colonies slow-growing, velvety, floccose; surface<br />

olivaceous-grey to dark olivaceous-green; reverse olivaceousblack.<br />

Hyphae smooth, thin-walled, pale olivaceous. Conidiophores<br />

arising vertically from creeping hyphae, straight or flexuose, simple,<br />

thick-walled, red-brown, with inflated basal cell. Conidiogenous<br />

cells terminally integrated, polyblastic, sympodial, smooth, thick-<br />

www.studiesinmycology.org<br />

89


Arzanlou et al.<br />

Fig. 34. Rhodoveronaea varioseptata (<strong>CBS</strong> 431.88). A–D. Macronematous conidiophores with sympodially proliferating conidiogenous cells, resulting in conidium bearing rachis<br />

with slightly prominent conidium-bearing denticles. E–F. Conidia with minute marginal frill. Scale bar = 10 µm.<br />

Fig. 35. Veronaeopsis simplex (<strong>CBS</strong> 588.66). A–C. Conidial apparatus at different stages of development, resulting in semi-micronematous conidiophores <strong>and</strong> sympodially<br />

proliferating conidiogenous cells. D–E. Rachis with crowded, prominent denticles. F. Intercalary conidiogenous cells. G. Conidia. Scale bar = 10 µm.<br />

walled, pale brown, rachis straight, occasionally geniculate, with<br />

crowded, slightly prominent conidium-bearing denticles; denticles<br />

flat-tipped, slightly pigmented. Conidia solitary, pale brown, thinor<br />

slightly thick-walled, smooth, ellipsoidal to obovoidal, 0–multiseptate,<br />

with a protruding base <strong>and</strong> a marginal basal frill; conidial<br />

secession schizolytic.<br />

Type species: Rhodoveronaea varioseptata Arzanlou, W. Gams &<br />

Crous, sp. nov.<br />

Notes: Rhodoveronaea differs from other ramichloridium-like fungi<br />

by the presence of a basal, marginal conidial frill, <strong>and</strong> variably<br />

septate conidia.<br />

90


Ramichloridium <strong>and</strong> allied genera<br />

Rhodoveronaea varioseptata Arzanlou, W. Gams & Crous, sp.<br />

nov. MycoBank MB504570. Figs 10D, 34.<br />

Etymology: Named for its variably septate conidia.<br />

Hyphae 2–3 µm latae. Conidiophora ad 125 µm longa et 3–5 µm lata. Cellulae<br />

conidiogenae 30–70 µm longae et 3–5 µm latae. Conidia 0–2(–3)-septata, (8–)11–<br />

13(–15) × (2–)3–4(–6) µm.<br />

In vitro: Submerged hyphae smooth, thin-walled, pale olivaceous,<br />

2–3 µm wide; aerial hyphae smooth, brownish <strong>and</strong> slightly narrower.<br />

Conidiophores arising vertically from creeping hyphae, straight or<br />

flexuose, simple, smooth, thick-walled, red-brown, up to 125 µm<br />

long, 3–5 µm wide, often with inflated basal cell. Conidiogenous<br />

cells terminally integrated, smooth, thick-walled, pale brown at the<br />

base, paler towards the apex, straight, variable in length, 30–70<br />

µm long <strong>and</strong> 3–5 µm wide, rachis straight, occasionally geniculate;<br />

slightly prominent conidium-bearing denticles, crowded, with slightly<br />

pigmented apex, about 1 µm diam. Conidia solitary, pale brown,<br />

thin- or slightly thick-walled, smooth, ellipsoid to obovoid, 0–2(–3)-<br />

septate, (8–)11–13(–15) × (2–)3–4(–6) µm with a protruding base,<br />

1.5 µm wide, <strong>and</strong> marginal frill.<br />

Cultural characteristics: Colonies reaching 12 mm diam after 14<br />

d at 24 °C, velvety, floccose; surface olivaceous-grey to dark<br />

olivaceous-green; reverse olivaceous-black.<br />

Specimen examined: Germany, Eifel, Berndorf, on Bertia moriformis, Sep. 1987, W.<br />

Gams, holotype <strong>CBS</strong>-H 19932, culture ex-type <strong>CBS</strong> 431.88.<br />

Venturiaceae (Pleosporales)<br />

<strong>The</strong> ex-type strain of Veronaea simplex (Papendorf 1969) did not<br />

cluster with the <strong>genus</strong> Veronaea (Herpotrichiellaceae), but is allied<br />

to the Venturiaceae. Veronaea simplex is distinct from species<br />

of Fusicladium Bonord. by having a well-developed rachis with<br />

densely aggregated scars. A new <strong>genus</strong> is thus introduced to<br />

accommodate this taxon.<br />

Veronaeopsis Arzanlou & Crous, gen. nov. MycoBank<br />

MB504571.<br />

Etymology: <strong>The</strong> suffix -opsis refers to its <strong>similar</strong>ity with Veronaea.<br />

Genus Veronaeae simile sed conidiophoris brevioribus (ad 60 μm longis) et rachide<br />

dense denticulata distinguendum.<br />

In vitro: Colonies moderately fast-growing; surface velvety, floccose,<br />

greyish sepia to hazel, with smooth margin; reverse mouse-grey<br />

to dark mouse-grey. Conidiophores arising vertically from aerial<br />

hyphae, lateral or intercalary, simple or branched, occasionally<br />

reduced to conidiogenous cells, pale brown. Conidiogenous<br />

cells terminally integrated on simple or branched conidiophores,<br />

polyblastic, smooth, thin-walled, pale brown; rachis commonly<br />

straight, geniculate, with densely crowded, prominent denticles,<br />

<strong>and</strong> slightly pigmented scars. Conidia solitary, subhyaline to pale<br />

brown, thin- or slightly thick-walled, smooth, oblong-ellipsoidal to<br />

subcylindrical, (0–)1-septate, with a slightly darkened, thickened,<br />

hilum; conidial secession schizolytic.<br />

Type species: Veronaeopsis simplex (Papendorf) Arzanlou &<br />

Crous, comb. nov.<br />

Veronaeopsis simplex (Papendorf) Arzanlou & Crous, comb.<br />

nov. MycoBank MB504572. Figs 17C, 35.<br />

Basionym: Veronaea simplex Papendorf, Trans. Brit. Mycol. Soc.<br />

52: 486. 1969.<br />

www.studiesinmycology.org<br />

In vitro: Submerged hyphae smooth, thin-walled, pale brown; aerial<br />

hyphae aggregated in bundles. Conidiophores arising vertically<br />

from aerial hyphae, lateral or intercalary, simple or branched,<br />

occasionally reduced to conidiogenous cells, pale brown, rather<br />

short, up to 60 µm long, 1.5–2 µm wide. Conidiogenous cells<br />

terminally integrated in the conidiophores, smooth, thin-walled, pale<br />

brown, variable in length, 5–25 µm long, rachis generally straight<br />

or irregularly geniculate, with crowded, prominent denticles, about<br />

0.5 µm long, flat-tipped, with slightly pigmented apex. Conidia<br />

solitary, subhyaline to pale brown, thin- or slightly thick-walled,<br />

smooth, oblong-ellipsoidal to subcylindrical, (0–)1-septate, slightly<br />

constricted at the septum, (6–)10–12(–15) × (2–)2.5–3(–4) µm;<br />

hilum slightly darkened <strong>and</strong> thickened, not refractive, about 1 µm<br />

diam.<br />

Cultural characteristics: Colonies reaching 25 mm diam after 14<br />

d at 24 °C; surface velvety, floccose, greyish sepia to hazel, with<br />

smooth margin; reverse mouse-grey to dark mouse-grey.<br />

Specimen examined: South Africa, Potchefstroom, on leaf litter of Acacia karroo,<br />

1966, M.C. Papendorf, holotype, <strong>CBS</strong> H-7810; culture ex-type <strong>CBS</strong> 588.66 = IMI<br />

203547.<br />

Notes: <strong>The</strong> presence of 1-septate conidia in Veronaeopsis overlaps<br />

with Veronaea. However, Veronaeopsis differs from Veronaea<br />

based on its conidiophore <strong>and</strong> conidiogenous cell morphology.<br />

Veronaea has much longer, macronematous conidiophores than<br />

Veronaeopsis. Furthermore, Veronaea has a more or less straight<br />

rachis, whereas in Veronaeopsis the rachis is often geniculate.<br />

<strong>The</strong> conidiogenous loci in Veronaea are less prominent, i.e., less<br />

denticle-like.<br />

Discussion<br />

<strong>The</strong> present study was initiated chiefly to clarify the status of<br />

Ramichloridium musae, the causal organism of tropical speckle<br />

disease of banana (Jones 2000). Much confusion surrounded this<br />

name in the past, relating, respectively, to its validation, species<br />

<strong>and</strong> generic status. As was revealed in the present study, however,<br />

two species are involved in banana speckle disease, namely R.<br />

musae <strong>and</strong> R. biverticillatum. Even more surprising was the fact<br />

that Ramichloridium comprises anamorphs of Mycosphaerella<br />

Johanson (Mycosphaerellaceae), though no teleomorphs have<br />

thus far been conclusively linked to any species of Ramichloridium.<br />

By investigating the Ramichloridium generic complex as outlined<br />

by de Hoog (1977), another <strong>genus</strong> associated with leaf spots,<br />

namely Periconiella, was also shown to represent an anamorph<br />

of Mycosphaerella. Although no teleomorph connections have<br />

been proven for ramichloridium-like taxa, de Hoog et al. (1983)<br />

refer to the type specimen of Wentiomyces javanicus Koord.<br />

(Pseudoperisporiaceae), on the type specimen of which (PC)<br />

some ramichloridium-like conidiophores were seen. Without<br />

fresh material <strong>and</strong> an anamorph-teleomorph connection proven<br />

in culture, however, this matter cannot be investigated further. It<br />

is interesting to note, however, that Wentiomyces Koord. shows<br />

a strong resemblance to Mycosphaerella, except for the external<br />

perithecial appendages.<br />

<strong>The</strong> <strong>genus</strong> Mycosphaerella is presently one of the largest genera<br />

of ascomycetes, containing close to 3 000 names (Aptroot 2006), to<br />

which approximately 30 anamorph genera have already been linked<br />

(Crous et al. 2006a, b, 2007). By adding two additional anamorph<br />

genera, the Mycosphaerella complex appears to be exp<strong>and</strong>ing<br />

91


Arzanlou et al.<br />

even further, though some taxa have been shown to reside in other<br />

families in the Capnodiales, such as Davidiella Crous & U. Braun<br />

(Davidiellaceae) <strong>and</strong> Teratosphaeria (Teratosphaeriaceae) (Braun<br />

et al. 2003, Crous et al. 2007, Schubert et al. 2007 – this volume).<br />

Another family, which proved to accommodate several<br />

ramichloridium-like taxa, is the Herpotrichiellaceae (Chaetothyriales).<br />

Members of the Chaetothyriales are regularly encountered as<br />

causal agents of human mycoses (Haase et al. 1999, de Hoog<br />

et al. 2003), whereas species of the Capnodiales are common<br />

plant pathogens, or chiefly associated with plants. Species in the<br />

Chaetothyriales have consistently melanized thalli, which is a factor<br />

enabling them to invade humans, <strong>and</strong> cause a wide diversity of<br />

mycoses, such as chromoblastomycosis, mycetoma, brain infection<br />

<strong>and</strong> subcutaneous phaeohyphomycosis (de Hoog et al. 2003). <strong>The</strong><br />

only known teleomorph connection in this <strong>genus</strong> is Capronia Sacc.<br />

(Untereiner & Naveau 1999).<br />

Rhinocladiella <strong>and</strong> Veronaea were in the past frequently<br />

confused with the <strong>genus</strong> Ramichloridium. However, Rhinocladiella,<br />

as well as Veronaea <strong>and</strong> Thysanorea, were shown to cluster in the<br />

Chaetothyriales, while Ramichloridium clusters in the Capnodiales.<br />

Rhinocladiella mackenziei, which causes severe cerebral<br />

phaeohyphomycosis in humans (Sutton et al. 1998), has in the<br />

past been confused with Pleurothecium obovoideum (Ur-Rahman<br />

et al. 1988). Data presented here reveal, however, that although<br />

morphologically <strong>similar</strong>, these species are phylogenetically<br />

separate, with P. obovoideum belonging to the Sordariales, where<br />

it clusters with sexual species of Carpoligna F.A. Fernández &<br />

Huhndorf that have Pleurothecium anamorphs (Fernández et al.<br />

1999).<br />

In addition to the genera clustering in the Capnodiales <strong>and</strong><br />

Chaetothyriales, several ramichloridium-like genera are newly<br />

introduced to accommodate species that cluster elsewhere<br />

in the ascomycetes, namely Pseudovirgaria, Radulidium <strong>and</strong><br />

Myrmecridium, Veronaeopsis, <strong>and</strong> Rhodoveronaea. Although the<br />

ecological role of these taxa is much less known than that of taxa in<br />

the Capnodiales <strong>and</strong> Chaetothyriales, some exhibit an interesting<br />

ecology. For instance, the fungicolous habit of Pseudovirgaria, as<br />

well as some species in Radulidium, which are found on various<br />

rust species, suggests that these genera should be screened<br />

further to establish if they have any potential biocontrol properties.<br />

Furthermore, these two genera share a common ancestor, <strong>and</strong><br />

further work is required to determine whether speciation was shaped<br />

by co-evolution with the rusts. A further species of “Veronaea” that<br />

might belong to Pseudovirgaria is Veronaea harunganae (Hansf.)<br />

M.B. Ellis, which is known to occur on Hemileia harunganae<br />

Cummins on Harungana in Tanzania <strong>and</strong> Ug<strong>and</strong>a (Ellis 1976). <strong>The</strong><br />

latter species, however, is presently not known from culture, <strong>and</strong><br />

needs to be recollected to facilitate further study.<br />

<strong>The</strong> genera distinguished here represent homogeneous clades<br />

in the phylogenetic analysis. Only the species of Rhinocladiella are<br />

dispersed among others morphologically classified in Exophiala or<br />

other genera.<br />

By integrating the phylogenetic data generated here with<br />

the various morphological data sets, we were able to resolve<br />

eight clades for taxa formerly regarded as representative of the<br />

Ramichloridium complex. According to the phylogeny inferred<br />

from 28S rDNA sequence data, the genera Ramichloridium <strong>and</strong><br />

Periconiella were heterogeneous, requiring the introduction of<br />

several novel genera. Although the present 11 odd genera can<br />

still be distinguished based on their morphology, it is unlikely that<br />

morphological identifications without the supplement of molecular<br />

data would in the future be able to accurately identify all the novel<br />

isolates that undoubtably await description. <strong>The</strong> integration of<br />

morphology with phylogenetic data not only helps to resolve generic<br />

affinities, but it also assists in discriminating between the various<br />

cryptic species that surround many of these well-known names that<br />

are presently freely used in the literature. To that end it is interesting<br />

to note that for the majority of the taxa studied here, the ITS domain<br />

(Table 1) provided good species resolution. However, more genes<br />

will have to be screened in future studies aimed at characterising<br />

some of the species complexes where the ITS domain provided<br />

insufficient phylogenetic signal (data not shown) to resolve all of the<br />

observed morphological species.<br />

ACKNOWLEDGEMENTS<br />

<strong>The</strong> work of Mahdi Arzanlou was funded by the Ministry of Science, Research <strong>and</strong><br />

Technology of Iran, which we gratefully acknowledge. Several colleagues from<br />

different countries provided material without which this work would not have been<br />

possible. We thank Marjan Vermaas for preparing the photographic plates, <strong>and</strong><br />

Arien van Iperen for taking care of the cultures.<br />

REFERENCES<br />

Al-Hedaithy SSA, Jamjoom ZAB, Saeed ES (1988). Cerebral phaeohyphomycosis<br />

caused by Fonsecaea pedrosoi in Saudi-Arabia. Acta Pathologica Microbiologica<br />

Sc<strong>and</strong>inaviae 96 Suppl. 3: 94–100.<br />

Aptroot A (2006). Mycosphaerella <strong>and</strong> its anamorphs: 2. Conspectus of<br />

Mycosphaerella. <strong>CBS</strong> Biodiversity Series 5: 1–231.<br />

Braun U (1998). A monograph of Cercosporella, Ramularia <strong>and</strong> allied genera<br />

(phytopathogenic hypomycetes). Vol. 2. IHW-Verlag, Eching.<br />

Braun U, Crous PW, Dugan F, Groenewald JZ, Hoog GS de (2003). Phylogeny <strong>and</strong><br />

taxonomy of cladosporium-like hyphomycetes, including Davidiella gen. nov.,<br />

the teleomorph of <strong>Cladosporium</strong> s. str. Mycological Progress 2: 3–18.<br />

Campbell CK, Al-Hedaithy SSA (1993). Phaeohyphomycosis of the brain caused<br />

by Ramichloridium mackenziei sp. nov. in middle eastern countries. Journal of<br />

Medical <strong>and</strong> Veterinary Mycology 31: 325–332.<br />

Crous PW, Groenewald JZ, Groenewald M, Caldwell P, Braun U, Harrington TC<br />

(2006a). Species of Cercospora associated with grey leaf spot of maize.<br />

Studies in Mycology 55: 189–197.<br />

Crous PW, Groenewald JZ, Risède J-M, Simoneau P, Hyde KD (2006b). Calonectria<br />

species <strong>and</strong> their Cylindrocladium anamorphs: species with clavate vesicles.<br />

Studies in Mycology 55: 213–226.<br />

Crous PW, Slippers B, Wingfield MJ, Rheeder J, Marasas WFO, Phillips AJL, Alves<br />

A, Burgess T, Barber P, Groenewald JZ (2006c). Phylogenetic lineages in the<br />

Botryosphaeriaceae. Studies in Mycology 55: 235–253.<br />

Crous PW, Summerell BA, Carnegie AJ, Mohammed C, Himaman W, Groenewald<br />

JZ (2007). Foliicolous Mycosphaerella spp. <strong>and</strong> their anamorphs on Corymbia<br />

<strong>and</strong> Eucalyptus. Fungal Diversity 26: 143–185.<br />

David JC (1997). A contribution to the systematics of <strong>Cladosporium</strong>. Revision of the<br />

fungi previously referred to Heterosporium. Mycological Papers 172: 1–157.<br />

Ellis MB (1971). Dematiaceous Hyphomycetes. Commonwealth Mycological<br />

Institute, Kew.<br />

Ellis MB (1976). More Dematiaceous Hyphomycetes. Commonwealth Mycological<br />

Institute, Kew.<br />

Fernández FA, Lutzoni FM, Huhndorf SM (1999). Teleomorph-anamorph<br />

connections: the new pyrenomycetous <strong>genus</strong> Carpoligna <strong>and</strong> its Pleurothecium<br />

anamorph. Mycologia 91: 251–262.<br />

Gams W, Verkleij GJM, Crous PW (eds) (2007). <strong>CBS</strong> Course of Mycology, 5 th ed.<br />

Centraalbureau voor Schimmelcultures, Utrecht.<br />

Haase G, Sonntag L, Melzer-Krick B, Hoog GS de (1999). Phylogenetic inference<br />

by SSU-gene analysis of members of the Herpotrichiellaceae with special<br />

reference to human pathogenic species. Studies in Mycology 47: 80–97.<br />

Hoog GS de (1977). Rhinocladiella <strong>and</strong> allied genera. Studies in Mycology 15:<br />

1–140.<br />

Hoog GS de (1979). Nomenclatural notes on some black yeast-like hyphomycetes.<br />

Taxon 28: 347–348.<br />

Hoog GS de, Rahman MA, Boekhout T (1983). Ramichloridium, Veronaea <strong>and</strong><br />

Stenella: generic delimitation, new combinations <strong>and</strong> two new species.<br />

Transactions of the British Mycological Society 81: 485–490.<br />

Hoog GS de, Vicente V, Caligiorne RB, Kantarcioglu S, Tintelnot KA, Gerrits van<br />

den Ende AHG, Haase G (2003). Species diversity <strong>and</strong> polymorphism in<br />

92


Ramichloridium <strong>and</strong> allied genera<br />

the Exophiala spinifera clade containing opportunistic black yeast-like fungi.<br />

Journal of Clinical Microbiology 41: 4767–4778.<br />

Hughes SJ (1951). Studies on Microfungi. 5. Acrotheca. Mycological Papers 38:<br />

1–8.<br />

Jones RD (2000). Tropical speckle. In: Disease of Banana, Abaca <strong>and</strong> Enset (Jones<br />

RD, ed.). CABI Publishing, Wallingford: 116–120.<br />

McGinnis MR, Schell WA (1980). <strong>The</strong> <strong>genus</strong> Fonsecaea <strong>and</strong> its relationship to the<br />

genera <strong>Cladosporium</strong>, Phialophora, Ramichloridium, <strong>and</strong> Rhinocladiella. Pan<br />

American Health Organisation Scientific Publication 396: 215–234.<br />

Morgan-Jones G (1979). Notes on hyphomycetes. 28. Veronaea bambusae sp. nov.<br />

Mycotaxon 8: 149–151.<br />

Morgan-Jones G (1982). Notes on Hyphomycetes. 40. New species of Codinaea<br />

<strong>and</strong> Veronaea. Mycotaxon 14: 175–180.<br />

Mostert L, Groenewald JZ, Summerbell RC, Gams W, Crous PW (2006). Taxonomy<br />

<strong>and</strong> pathology of Togninia (Diaporthales) <strong>and</strong> its Phaeoacremonium anamorphs.<br />

Studies in Mycology 54: 1–113.<br />

Nyl<strong>and</strong>er JAA (2004). MrModeltest v2.2. Program distributed by the author.<br />

Evolutionary Biology Centre, Uppsala University.<br />

Papendorf MC (1969). New South African soil Fungi. Transactions of the British<br />

Mycological Society 52: 483–489.<br />

Papendorf MC (1976). Notes on Veronaea including V. compacta sp. nov. Bothalia<br />

12: 119–121.<br />

Rayner RW (1970). A mycological colour chart. CMI <strong>and</strong> British Mycological Society.<br />

Kew.<br />

Rehner SA, Samuels GJ (1994). Taxonomy <strong>and</strong> phylogeny of Gliocladium analysed<br />

from nuclear large subunit ribosomal DNA sequences. Mycological Research<br />

98: 625–634.<br />

Ronquist F, Huelsenbeck JP (2003). MrBayes 3: Bayesian phylogenetic inference<br />

under mixed models. Bioinformatics 19: 1572–1574.<br />

Saccardo PA, Berlese AN (1885). Miscellanea mycologica. Ser. II. Atti dell’Istituto<br />

Veneto di Scienze, Lettere ed Arti Ser. 6 3: 711–742.<br />

Schol-Schwarz MB (1968). Rhinocladiella, its synonym Fonsecaea <strong>and</strong> its relation<br />

to Phialophora. Antonie van Leeuwenhoek 34: 119–152.<br />

Schubert K, Groenewald JZ, Braun U, Dijksterhuis J, Starink M, Hill CF, Zalar P,<br />

Hoog GS de, Crous PW (2007). Biodiversity in the <strong>Cladosporium</strong> herbarum<br />

complex (Davidiellaceae, Capnodiales), with st<strong>and</strong>ardisation of methods for<br />

<strong>Cladosporium</strong> taxonomy <strong>and</strong> diagnostics. Studies in Mycology 58: 105–156.<br />

Seifert KA (1993). Integrating anamorphic fungi into the fungal system. In: <strong>The</strong><br />

Fungal Holomorph: Mitotic, Meiotic <strong>and</strong> Pleomorphic Speciation in Fungal<br />

Systematics (Reynolds DR, Taylor JW, eds). International Mycological Institute,<br />

Egham: 79–85.<br />

Stahel G (1937). <strong>The</strong> banana leaf speckle in Surinam caused by Chloridium musae<br />

nov. spec. <strong>and</strong> another related banana disease. Tropical Agriculture 14: 42–<br />

44.<br />

Sutton DA, Slifkin M, Yakulis R, Rinaldi MG (1998). U.S. case report of cerebral<br />

phaeohyphomycosis caused by Ramichloridium obovoideum (R. mackenziei):<br />

Criteria for identification, therapy, <strong>and</strong> review of other known <strong>dematiaceous</strong><br />

neurotropic taxa. Journal of Clinical Microbiology 36: 708–715.<br />

Swofford DL (2003). PAUP*. Phylogenetic Analysis Using Parsimony (*<strong>and</strong> other<br />

Methods). Version 4. Sinauer Associates, Sunderl<strong>and</strong>, Massachusetts.<br />

Taylor JW, Jacobson DJ, Kroken S, Kasuga T, Geiser DM, Hibbett DS, Fisher MC<br />

(2000). Phylogenetic species recognition <strong>and</strong> species concepts in fungi. Fungal<br />

Genetics <strong>and</strong> Biology 31: 21–32.<br />

Untereiner WA, Naveau FA (1999). Molecular systematics of the Herpotrichiellaceae<br />

with an assessment of the phylogenetic positions of Exophiala dermatitidis <strong>and</strong><br />

Phialophora americana. Mycologia 91: 67–83.<br />

Ur-Rahman N, Mahgoub E, Chagla AH (1988). Fatal brain abscesses caused by<br />

Ramichloridium obovoideum – Report of three cases. Acta Neurochirurgica 93:<br />

92–95.<br />

Veerkamp J, Gams W (1983). Los hongos de Colombia – VIII. Some new species of<br />

soil fungi from Colombia. Caldasia 13: 709–717.<br />

Vilgalys R, Hester M (1990). Rapid genetic identification <strong>and</strong> mapping of<br />

enzymatically amplified ribosomal DNA from several Cryptococcus species.<br />

Journal of Bacteriology 172: 4238–4246.<br />

White TJ, Bruns T, Lee S, Taylor J (1990). Amplification <strong>and</strong> direct sequencing of<br />

fungal ribosomal RNA genes for phylogenetics. In: PCR Protocols: a guide to<br />

methods <strong>and</strong> applications (Innis MA, Gelf<strong>and</strong> DH, Sninsky JJ, White TJ, eds).<br />

Academic Press, San Diego, California: 315–322.<br />

Zipfel RD, Beer W de, Jacobs K, Wingfield BD, Wingfield MJ (2006). Multi-gene<br />

phylogenies define Ceratocystiopsis <strong>and</strong> Grosmannia distinct from Ophiostoma.<br />

Studies in Mycology 55: 75–97.<br />

www.studiesinmycology.org<br />

93


available online at www.studiesinmycology.org<br />

doi:10.3114/sim.2007.58.04<br />

Studies in Mycology 58: 95–104. 2007.<br />

<strong>Cladosporium</strong> leaf-blotch <strong>and</strong> stem rot of Paeonia spp. caused by Dichocladosporium<br />

chlorocephalum gen. nov.<br />

K. Schubert 1* , U. Braun 2 , J.Z. Groenewald 3 <strong>and</strong> P.W. Crous 3<br />

1<br />

Botanische Staatssammlung München, Menzinger Strasse 67, D-80638 München, Germany; 2 Martin-Luther-Universität, Institut für Biologie, Geobotanik und Botanischer Garten,<br />

Herbarium, Neuwerk 21, D-06099 Halle (Saale), Germany; 3 <strong>CBS</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD Utrecht, <strong>The</strong> Netherl<strong>and</strong>s<br />

*Correspondence: Konstanze Schubert, konstanze.schubert@gmx.de<br />

Abstract: <strong>Cladosporium</strong> chlorocephalum (= C. paeoniae) is a common, widespread leaf-spotting hyphomycete of peony (Paeonia spp.), characterised by having dimorphic<br />

conidiophores. During the season, one stage of this fungus causes distinct, necrotic leaf-blotch symptoms on living leaves of Paeonia spp. In late autumn, winter or after<br />

overwintering, a second morphologically distinct conidiophore type occurs on dead, blackish, rotting stems. Conspecificity of the two morphs, previously proposed on the basis<br />

of observations in culture, was supported by DNA sequence data from the ITS <strong>and</strong> LSU gene regions, using cultures obtained from leaf-blotch symptoms on living leaves, as<br />

well as from dead stems of Paeonia spp. Sequence data were identical, indicating a single species with two morphs. On account of its distinct conidiogenous loci <strong>and</strong> conidial<br />

hila, as well as its sequence-based phylogenetic position separate from the Davidiella/<strong>Cladosporium</strong> clade, the peony fungus has to be excluded from <strong>Cladosporium</strong> s. str., but<br />

still belongs to the Davidiellaceae (Capnodiales). <strong>The</strong> leaf-blotching (cladosporioid) morph of this fungus morphologically resembles species of Fusicladium, but differs in having<br />

dimorphic fruiting, <strong>and</strong> is phylogenetically distant from the Venturiaceae. <strong>The</strong> macronematous (periconioid) morph resembles Metulocladosporiella (Chaetothyriales), but lacks<br />

rhizoid conidiophore hyphae, <strong>and</strong> has 0–5-septate conidia. Hence, C. chlorocephalum is assigned to the new <strong>genus</strong> Dichocladosporium.<br />

Taxonomic novelties: Dichocladosporium K. Schub., U. Braun & Crous, gen. nov., Dichocladosporium chlorocephalum (Fresen.) K. Schub., U. Braun & Crous, comb. nov.<br />

Key words: Anamorphic fungi, <strong>Cladosporium</strong> chlorocephalum, C. paeoniae, hyphomycetes, new <strong>genus</strong>, phylogeny, taxonomy.<br />

Introduction<br />

Fresenius (1850) described Periconia chlorocephala Fresen. from<br />

Germany on dead stems of Paeonia sp. Mason & Ellis (1953)<br />

examined this species in vitro <strong>and</strong> in vivo <strong>and</strong> stated that it only<br />

occurred on dead stems of Paeonia spp. <strong>The</strong>y described, illustrated<br />

<strong>and</strong> discussed this species in detail, <strong>and</strong> reallocated it to the <strong>genus</strong><br />

<strong>Cladosporium</strong> Link.<br />

A second, cladosporioid hyphomycete on Paeonia spp.,<br />

<strong>Cladosporium</strong> paeoniae Pass., was collected by Passerini on living<br />

leaves of P. albiflora (as P. edulis) in Italy, <strong>and</strong> distributed in Thümen,<br />

Herbarium mycologicum oeconomicum, Fasc. IX, No. 416 (1876),<br />

together with the first valid description, which was repeated by<br />

Passerini (1876). Later, Passerini collected this fungus on Paeonia<br />

officinalis at Parma in Italy <strong>and</strong> distributed it in Thümen, Mycotheca<br />

universalis, No. 670 (1877). Saccardo (1882) listed a collection of<br />

this species on Paeonia anomala from Russia, Siberia, which he<br />

later described as <strong>Cladosporium</strong> paeoniae var. paeoniae-anomalae<br />

Sacc. (Saccardo 1886). A first examination of C. paeoniae in culture<br />

was accomplished by Meuli (1937), followed by a treatment in vitro<br />

by de Vries (1952). Mason & Ellis (1953) described <strong>and</strong> illustrated in<br />

their treatment of C. chlorocephalum macroconidiophores, agreeing<br />

with those of the original diagnosis <strong>and</strong> illustration of Periconia<br />

chlorocephala, as well as semi-macronematous conidiophores<br />

concurring with those of C. paeoniae, although no mention was<br />

made of the latter name. McKemy & Morgan-Jones (1991) carried<br />

out comprehensive studies on <strong>Cladosporium</strong> on Paeonia spp. in<br />

vitro <strong>and</strong> in vivo, including detailed discussions of the history of<br />

the taxa concerned, taxonomic implications <strong>and</strong> comprehensive<br />

descriptions <strong>and</strong> illustrations. <strong>The</strong>y concluded that <strong>Cladosporium</strong><br />

paeoniae, found in culture together with C. chlorocephalum, was<br />

a semi-macronematous form (synanamorph) of the latter species,<br />

<strong>and</strong> reduced C. paeoniae to synonymy with the latter species.<br />

In the present study, re-examination <strong>and</strong> reassessment of<br />

morphological characters, conidiogenesis, <strong>and</strong> DNA sequence<br />

data of the ITS <strong>and</strong> 28S nrDNA were used to confirm the identity<br />

of <strong>Cladosporium</strong> chlorocephalum (the periconioid morph) <strong>and</strong> C.<br />

paeoniae (the cladosporioid morph), <strong>and</strong> clarify their relation to<br />

<strong>Cladosporium</strong> s. str. (Davidiellaceae) (emend. David 1997, Braun<br />

et al. 2003).<br />

Materials <strong>and</strong> methods<br />

Isolates<br />

Single-conidial isolates were obtained from symptomatic leaves<br />

<strong>and</strong> dead stems, <strong>and</strong> cultured as detailed in Crous (1998). Cultural<br />

characteristics <strong>and</strong> morphology of isolates (Table 1) were recorded<br />

from plates containing either 2 % potato-dextrose agar (PDA) or<br />

synthetic nutrient-poor agar (SNA) (Gams et al. 2007). Plates were<br />

incubated at 25 °C under continuous near-UV light to promote<br />

sporulation.<br />

DNA isolation, amplification <strong>and</strong> sequencing<br />

Fungal colonies were established on agar plates, <strong>and</strong> genomic<br />

DNA was isolated following the protocol of Lee & Taylor (1990).<br />

<strong>The</strong> primers V9G (de Hoog & Gerrits van den Ende 1998) <strong>and</strong> LR5<br />

(Vilgalys & Hester 1990) were used to amplify part (ITS) of the nuclear<br />

rDNA operon spanning the 3’ end of the 18S rRNA gene (SSU), the<br />

first internal transcribed spacer (ITS1), the 5.8S rRNA gene, the<br />

second ITS region <strong>and</strong> the 5’ end of the 28S rRNA gene (LSU).<br />

<strong>The</strong> primers ITS4 (White et al. 1990), LR0R (Rehner & Samuels<br />

1994), LR3R (www.biology.duke.edu/fungi/mycolab/primers.htm)<br />

<strong>and</strong> LR16 (Moncalvo et al. 1993) were used as internal sequence<br />

primers to ensure good quality sequences over the entire length<br />

of the amplicon. <strong>The</strong> PCR conditions, sequence alignment <strong>and</strong><br />

subsequent phylogenetic analysis followed the methods of Crous<br />

et al. (2006b). Gaps longer than 10 bases were coded as single<br />

events for the phylogenetic analyses; the remaining gaps were<br />

treated as new character states. Sequence data were deposited<br />

in GenBank (Table 1) <strong>and</strong> alignments in TreeBASE (www.treebase.<br />

org).<br />

95


Schubert et al.<br />

Morphology<br />

Morphological examinations were made from herbarium samples,<br />

fresh symptomatic leaves <strong>and</strong> stems, as well as cultures sporulating<br />

on SNA. Structures were mounted in water or Shear’s solution<br />

(Dhingra & Sinclair 1985), <strong>and</strong> 30 measurements at × 1 000<br />

magnification were made of each structure under an Olympus BX<br />

50 microscope (Hamburg, Germany). <strong>The</strong> 95 % confidence levels<br />

were determined <strong>and</strong> the extremes of spore measurements given<br />

in parentheses. Scanning electron microscopic examinations were<br />

conducted at the Institute of Zoology, Martin-Luther-University,<br />

Halle (Saale), Germany, using a Hitachi S-2400. Samples were<br />

coated with a thin layer of gold applied with a sputter coater SCD<br />

004 (200 s in an argon atmosphere of 20 mA, 30 mm distant<br />

from the electrode). Colony colours were noted after 2 wk growth<br />

on PDA at 25 °C in the dark, using the colour charts of Rayner<br />

(1970). All cultures studied were deposited in the culture collection<br />

of the Centraalbureau voor Schimmelcultures (<strong>CBS</strong>), Utrecht,<br />

the Netherl<strong>and</strong>s (Table 1). Taxonomic novelties were lodged with<br />

MycoBank (www.MycoBank.org).<br />

Results<br />

DNA phylogeny<br />

Amplification products of approximately 1 700 bases were obtained<br />

for the isolates listed in Table 1. <strong>The</strong> ITS region of the sequences<br />

was used to obtain additional sequences from GenBank which were<br />

added to the alignment. <strong>The</strong> manually adjusted alignment contained<br />

26 sequences (including the two outgroup sequences) <strong>and</strong> 518<br />

characters including alignment gaps. Of the 518 characters used<br />

in the phylogenetic analysis, 226 were parsimony-informative, 33<br />

were variable <strong>and</strong> parsimony-uninformative, <strong>and</strong> 259 were constant.<br />

Neighbour-joining analysis using three substitution models on the<br />

sequence data yielded trees supporting the same clades but with<br />

a different arrangement at the deeper nodes. <strong>The</strong>se nodes were<br />

supported poorly in the bootstrap analyses (the highest value<br />

observed for one of these nodes was 64 %; data not shown).<br />

Two equally most parsimonious trees (TL = 585 steps; CI =<br />

0.761; RI = 0.902; RC = 0.686), one of which is shown in Fig. 1,<br />

was obtained from the parsimony analysis of the ITS region. <strong>The</strong><br />

Dichocladosporium K. Schub., U. Braun & Crous isolates formed a<br />

well-supported clade (100 % bootstrap support), distinct from clades<br />

containing species of Davidiella Crous & U. Braun, Mycosphaerella<br />

Johanson <strong>and</strong> Teratosphaeria Syd. & P. Syd. This placement was<br />

also supported by analyses of the first part of the 28S rRNA gene<br />

(see Crous et al. 2007 – this volume).<br />

Taxonomy<br />

Because conidia formed holoblastically in simple or branched<br />

acropetal chains, <strong>Cladosporium</strong> chlorocephalum <strong>and</strong> C. paeoniae<br />

coincided with previous concepts of <strong>Cladosporium</strong> s. lat. (Braun<br />

et al. 2003, Schubert 2005), belonging to a wide assemblage of<br />

genera classified by Kiffer & Morelet (1999) as “Acroblastosporae”.<br />

Previous studies conducted in vitro concluded that <strong>Cladosporium</strong><br />

chlorocephalum <strong>and</strong> C. paeoniae represent two developmental<br />

stages (morphs) of a single species, a result confirmed here by<br />

DNA sequence analyses. A detailed analysis of conidiogenesis,<br />

structure of the conidiogenous loci <strong>and</strong> conidial hila, <strong>and</strong> a<br />

comparison with <strong>Cladosporium</strong> s. str., typified by C. herbarum<br />

(Pers. : Fr.) Link, revealed obvious differences: <strong>The</strong> conidiogenous<br />

loci <strong>and</strong> conidial hila of C. chlorocephalum are quite distinct from<br />

those of <strong>Cladosporium</strong> s. str. by being denticulate or subdenticulate,<br />

apically broadly truncate, unthickened or slightly thickened, but<br />

somewhat darkened-refractive. <strong>The</strong> scars in <strong>Cladosporium</strong> s.<br />

str. are, however, characteristically coronate, i.e., with a central<br />

convex dome surrounded by a raised periclinal rim (David 1997,<br />

Braun et al. 2003, Schubert 2005). Hence, the peony fungus<br />

has to be excluded from <strong>Cladosporium</strong> s. str. A comparison with<br />

phaeoblastic hyphomycetous genera revealed a close <strong>similar</strong>ity of<br />

this fungus with the <strong>genus</strong> Metulocladosporiella Crous, Schroers,<br />

J.Z. Groenew., U. Braun & K. Schub. recently introduced for the<br />

<strong>Cladosporium</strong> speckle disease of banana (Crous et al. 2006a).<br />

Both fungi have dimorphic fruiting, pigmented macronematous<br />

conidiophores often with distinct basal swellings <strong>and</strong> densely<br />

branched terminal heads composed of short branchlets <strong>and</strong><br />

ramoconidia, denticulate or subdenticulate unthickened, but<br />

somewhat darkened-refractive conidiogenous loci, as well as<br />

phaeoblastic conidia, formed in simple or branched acropetal<br />

chains. <strong>The</strong> semi-macronematous leaf-blotching morph is close<br />

to <strong>and</strong> barely distinguishable from Fusicladium Bonord. However,<br />

unlike Metulocladosporiella, the peony fungus does not form rhizoid<br />

hyphae at the base of conidiophore swellings <strong>and</strong> the conidia are<br />

amero- to phragmosporous [0–5-septate versus 0(–1)-septate<br />

in Metulocladosporiella]. Furthermore, the peony fungus neither<br />

clusters within the Chaetothyriales (with Metulocladosporiella)<br />

nor within the Venturiaceae (with Fusicladium), but clusters basal<br />

to the Davidiellaceae (see also Crous et al. 2007 – this volume).<br />

Hence, we propose to place C. chlorocephalum in the new <strong>genus</strong><br />

Dichocladosporium.<br />

Dichocladosporium K. Schub., U. Braun & Crous, gen. nov.<br />

MycoBank MB504428. Figs 2–5.<br />

Etymology: dicha in Greek = twofold.<br />

Differt a Metulocladosporiella conidiophoris cum cellulis basalibus saepe inflatis,<br />

sed sine hyphis rhizoidibus, conidiis amero- ad phragmosporis (0–5-septatis).<br />

Type species: Dichocladosporium chlorocephalum (Fresen.) K.<br />

Schub., U. Braun & Crous, comb. nov.<br />

Dichocladosporium chlorocephalum (Fresen.) K. Schub., U.<br />

Braun & Crous, comb. nov. MycoBank MB504429. Figs 2–5.<br />

Basionym: Periconia chlorocephala Fresen., Beiträge zur Mykologie<br />

1: 21. 1850.<br />

≡ Haplographium chlorocephalum (Fresen.) Grove, Sci. Gossip 21: 198.<br />

1885.<br />

≡ Graphiopsis chlorocephala (Fresen.) Trail, Scott. Naturalist (Perth) 10:<br />

75. 1889.<br />

≡ <strong>Cladosporium</strong> chlorocephalum (Fresen.) E.W. Mason & M.B. Ellis,<br />

Mycol. Pap. 56: 123. 1953.<br />

= <strong>Cladosporium</strong> paeoniae Pass., in Thümen, Herb. Mycol. Oecon., Fasc. IX,<br />

No. 416. 1876, <strong>and</strong> in Just´s Bot. Jahresber. 4: 235. 1876.<br />

= Periconia ellipsospora Penz. & Sacc., Atti Reale Ist. Veneto Sci. Lett. Arti,<br />

Ser. 6, 2: 596. 1884.<br />

= <strong>Cladosporium</strong> paeoniae var. paeoniae-anomalae Sacc., Syll. Fung. 4: 351.<br />

1886.<br />

= Haplographium chlorocephalum var. ovalisporum Ferraris, Fl. Ital. Cryptog.,<br />

Hyphales: 875. 1914.<br />

Descriptions: Mason & Ellis (1953: 123–126), McKemy & Morgan-<br />

Jones (1991: 140–144), Schubert (2005: 216).<br />

Illustrations: Fresenius (1850: Pl. IV, figs 10–15), Mason & Ellis<br />

(1953: 124–125, figs 42–43), McKemy & Morgan-Jones (1991:<br />

137, fig. 1; 141, fig. 2; 139, pl. 1; 143, pl. 2), Schubert (2005: 217,<br />

fig. 113; 275, pl. 34).<br />

96


Dichocladosporium gen. nov.<br />

10 changes<br />

Botryosphaeria ribis DQ316076<br />

Botryosphaeria stevensii AY343484<br />

100<br />

97<br />

70<br />

100<br />

100<br />

100<br />

97<br />

100<br />

100<br />

69<br />

<strong>Cladosporium</strong> tenellum CPC 12053<br />

<strong>Cladosporium</strong> variabile CPC 12753<br />

<strong>Cladosporium</strong> herbarum CPC 11600<br />

<strong>Cladosporium</strong> macrocarpum <strong>CBS</strong> 299.67<br />

<strong>Cladosporium</strong> bruhnei CPC 12211<br />

<strong>Cladosporium</strong> ossifragi <strong>CBS</strong> 842.91<br />

<strong>Cladosporium</strong> iridis <strong>CBS</strong> 138.40<br />

CPC 11383<br />

CPC 11969<br />

<strong>CBS</strong> 213.73<br />

<strong>CBS</strong> 100405<br />

100<br />

91<br />

Mycosphaerella punctiformis AY490763<br />

Passalora arachidicola AF297224<br />

Mycosphaerella aurantia AY509744<br />

Passalora fulva AF393701<br />

Cercospora apii AY840512<br />

Cercospora beticola AY840527<br />

Readeriella mirabilis AY725529<br />

Readeriella novaezel<strong>and</strong>iae DQ267603<br />

Teratosphaeria readeriellophora AY725577<br />

Teratosphaeria microspora AY260098<br />

Trimmatostroma abietis AY559362<br />

Trimmatostroma abietina AJ244267<br />

Trimmatostroma salicis AJ244264<br />

Dichocladosporium chlorocephalum<br />

Fig. 1. One of two equally most parsimonious trees obtained from a heuristic search with 100 r<strong>and</strong>om taxon additions of the ITS sequence alignment. <strong>The</strong> scale bar shows<br />

10 changes, <strong>and</strong> bootstrap support values from 1 000 replicates are shown at the nodes. Thickened lines indicate the strict consensus branches. <strong>The</strong> tree was rooted to two<br />

Botryosphaeria species.<br />

Characters of the cladosporioid morph: Leaf-blotch symptoms on<br />

living leaves amphigenous, variable in shape <strong>and</strong> size, subcircularoval<br />

to irregular, broad, oblong to exp<strong>and</strong>ed, up to 30 mm long<br />

<strong>and</strong> 20 mm wide, at times covering the entire leaf surface, forming<br />

olivaceous-brown to blackish brown patches, rarely violet-brown,<br />

margin usually indefinite, attacked areas turning dry with age,<br />

also occurring on young, green stems. Colonies amphigenous,<br />

punctiform to effuse, loose to dense, caespitose, brown, villose.<br />

Mycelium immersed, subcuticular to intraepidermal; hyphae<br />

sparingly branched, 4–7(–10) µm wide, septate, sometimes with<br />

swellings <strong>and</strong> constrictions, swollen cells up to 13 µm diam,<br />

subhyaline to pale brown, smooth, walls thickened, hyphae<br />

sometimes aggregated; in vitro mycelium at first mainly immersed,<br />

later also superficial, branched, 1–5(–7) µm wide, pluriseptate,<br />

often constricted at septa <strong>and</strong> with swellings <strong>and</strong> constrictions,<br />

therefore irregular in outline, smooth to verruculose or irregularly<br />

rough-walled, loosely verruculose with distinct large warts. Semimacronematous<br />

conidiophores formed on leaf-blotches solitary<br />

or in small, loose groups, arising from internal hyphae or swollen<br />

hyphal cells, erumpent through the cuticle, occasionally emerging<br />

through stomata, erect, straight to somewhat flexuous, oblongcylindrical,<br />

usually unbranched or occasionally branched, 13–80<br />

(–120) × (4–)5–8(–10) µm, slightly attenuated towards the apex,<br />

septate, septa often dense, unconstricted, pale to medium brown,<br />

sometimes paler towards the apex, smooth, thick-walled, wall often<br />

with two distinct layers, often somewhat inflated at the very base,<br />

up to 14 µm diam, occasionally proliferating enteroblastically; in<br />

vitro conidiophores arising laterally from plagiotropous hyphae or<br />

terminally from ascending hyphae, the latter usually appearing<br />

more filiform than those arising laterally from plagiotropous hyphae,<br />

erect, straight to slightly flexuous, cylindrical-oblong, not geniculate,<br />

usually unbranched, rarely with a short lateral prolongation near<br />

the apex, 18–60(–100) × 3–6 µm, slightly attenuated towards<br />

the apex, septate, pale to medium brown or olivaceous-brown,<br />

smooth to asperulate, walls somewhat thickened. Conidiogenous<br />

cells integrated, terminal or intercalary, subcylindrical, 7–45 µm<br />

www.studiesinmycology.org<br />

97


Schubert et al.<br />

Fig. 2. Dichocladosporium chlorocephalum (HAL 1924 F), periconioid, stem rotting morph. Conidiophores <strong>and</strong> conidia. Scale bar = 10 µm.<br />

Table 1. Isolates subjected to DNA analysis <strong>and</strong> morphological examination.<br />

Species Accession number 1 Host Country Collector GenBank accession<br />

number<br />

Dichocladosporium chlorocephalum <strong>CBS</strong> 213.73; IMI 048108a Paeonia sp. United Kingdom F. Rilstone EU009455<br />

<strong>CBS</strong> 100405 Paeonia sp. New Zeal<strong>and</strong> M. Braithwaite EU009456<br />

<strong>CBS</strong> 121522; CPC 11383 Paeonia delavayi Germany K. Schubert EU009457<br />

<strong>CBS</strong> 121523*; CPC 11969 Paeonia officinalis Germany K. Schubert EU009458<br />

1<br />

<strong>CBS</strong>: Centraalbureau voor Schimmelcultures, Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; CPC: Culture collection of Pedro Crous, housed at <strong>CBS</strong>; IMI: International<br />

Mycological Institute, CABI-Bioscience, Egham, Bakeham Lane, U.K.<br />

*Ex-type cultures.<br />

98


Dichocladosporium gen. nov.<br />

Fig. 3. Dichocladosporium chlorocephalum (<strong>CBS</strong> 121523 = CPC 11969). A–B. Symptoms of the periconioid, stem rotting morph. C–E, H. Macroconidiophores <strong>and</strong> conidia. F–G.<br />

Semi-macronematous conidiophores <strong>and</strong> conidia. I–J. Ramoconidia <strong>and</strong> conidia. K–M. Scanning electron microscopic photographs. K. Conidiophores. L. Conidial chain. M.<br />

Single conidium showing the surface ornamentation <strong>and</strong> scar structure. Scale bars: C–J, L = 10 µm; K = 100 µm; M = 2 µm.<br />

www.studiesinmycology.org<br />

99


Schubert et al.<br />

Fig. 4. Dichocladosporium chlorocephalum (HAL 2011 F), cladosporioid, leaf-spotting morph. Conidiophores <strong>and</strong> conidia. Scale bar = 10 µm.<br />

long, proliferation sympodial, with one to several conidiogenous<br />

loci, subdenticulate or denticulate, protuberant, terminally broadly<br />

truncate, 1.5–3 µm wide, unthickened or almost so, somewhat<br />

darkened-refractive. Conidia catenate, in simple or branched<br />

chains, polymorphous, small conidia globose, subglobose, broadly<br />

obovoid, 3–9 × 3–5 µm, aseptate, pale to medium brown, smooth,<br />

intercalary conidia limoniform, ellipsoid-fusiform, oblong, 5–23<br />

× 3.5–6.5 µm, 0–2-septate, medium brown, smooth to minutely<br />

verruculose or irregularly rough-walled, large conidia ellipsoid,<br />

oblong-cylindrical, ampulliform, 22–45(–52) × (4.5–)5–8 µm,<br />

0–5-septate, medium brown, smooth to minutely verruculose or<br />

irregularly rough-walled, walls somewhat thickened, hila truncate,<br />

1–3 µm wide, unthickened or almost so, somewhat darkenedrefractive;<br />

occasionally with microcyclic conidiogenesis; in vitro<br />

numerous, polymorphous, catenate, in loosely branched chains,<br />

small conidia globose, subglobose, or obovoid, 3–8 × 3–4 µm,<br />

aseptate, intercalary conidia limoniform to ellipsoid-fusiform, 9–18<br />

× 3.5–4.5 µm, 0–1-septate, large conidia ellipsoid to cylindricaloblong,<br />

14–30(–38) × 3–6(–7) µm, 0–3-septate, pale to medium<br />

brown, asperulate, minutely verruculose to irregularly roughwalled,<br />

walls thickened, hila usually short denticle-like, protuberant,<br />

truncate, in smaller conidia 0.5–1.8 µm wide, in larger conidia<br />

(1.5–)2–3 µm wide, unthickened or almost so but usually darkenedrefractive;<br />

with occasional microcyclic conidiogenesis.<br />

Characters of the periconioid morph: Macronematous conidiophores<br />

formed on faded or dead stems in late autumn, winter or after<br />

overwintering; colonies at first visible as reddish brown streaks, later<br />

turning olivaceous-brown to black, sometimes linear, sometimes<br />

encircling the stems, often occupying large stem segments, effuse,<br />

densely caespitose, velvety. Mycelium immersed, subcuticular<br />

100


Dichocladosporium gen. nov.<br />

to intraepidermal; hyphae at first sparsely branched, 3–7 µm<br />

wide, septate, not constricted at the septa, becoming swollen<br />

<strong>and</strong> wider, up to 11 µm wide, often branched, pale to medium<br />

olivaceous-brown, walls thickened, forming loose to dense hyphal<br />

aggregations; in vitro mycelium immersed to superficial, loosely<br />

branched, 2–6(–7) µm wide, pluriseptate, usually without swellings<br />

<strong>and</strong> constrictions, subhyaline to medium brown or olivaceousbrown,<br />

almost smooth to asperulate or irregularly rough-walled, in<br />

older colonies on PDA up to 10 µm wide, sometimes single hyphal<br />

cells distinctly swollen, up to 16(–20) µm wide, mainly at the base<br />

of conidiophores, sometimes covered by a slime coat or enveloped<br />

in a polysaccharide-like layer. Stromata well-developed, large <strong>and</strong><br />

exp<strong>and</strong>ed, up to about 50–320 µm in length, 15–30 µm deep,<br />

composed of a single to several layers of swollen pale to medium<br />

brown stromatic cells, 5–18 µm diam, thick-walled. Conidiophores<br />

solitary or in loose groups, arising from swollen hyphal cells or<br />

stromata, erumpent through the cuticle, erect, straight, rigid to<br />

slightly flexuous, 150–680 µm long, composed of a subcylindrical<br />

stipe, 13–24 µm wide at the base, slightly attenuated towards<br />

the apex, 5–15 µm just below the branched head, pluriseptate,<br />

not constricted at the septa, young conidiophores pale medium<br />

olivaceous-brown, later medium to usually dark brown, sometimes<br />

slightly paler at the distal end, smooth or almost so, often appearing<br />

somewhat granular, roughened, walls distinctly thickened, 1.5–3(–<br />

4) µm wide; apex with a roughly subglobose to ovoid head, about<br />

35–70 µm diam, composed of dense branchlets <strong>and</strong> ramoconidia,<br />

primary branchlets close to the apex <strong>and</strong> below the first <strong>and</strong><br />

sometimes second <strong>and</strong> third septa, solitary, in pairs or small verticils,<br />

appressed against the stipe or somewhat divergent, subcylindrical<br />

to ellipsoid-oval, aseptate, rarely 1-septate, pale olivaceous to dark<br />

brown, 10–20 × 5–8.5 µm; in vitro conidiophores initially micro<strong>and</strong><br />

semimacronematous, then progressively macronematous<br />

as colonies age, arising laterally from plagiotropous hyphae or<br />

terminally from ascending hyphae, sometimes also from swollen<br />

hyphal cells; micronematous conidiophores filiform, narrowly<br />

cylindrical-oblong, unbranched, up to 150 µm long, 2–3.5 µm wide,<br />

septate, septa often appear to be darkened, pale to pale medium<br />

olivaceous-brown, asperulate, walls slightly thickened; semimacronematous<br />

conidiophores often resembling those formed<br />

by the leaf-blotching (cladosporioid) morph on the natural host,<br />

subcylindrical to cylindrical-oblong, straight to slightly flexuous,<br />

unbranched, rarely branched, (10–)15–120 × 3–5(–6) µm, slightly<br />

attenuated towards the apex, septate, medium brown, minutely<br />

verruculose to irregularly rough-walled, walls more or less thickened;<br />

macronematous conidiophores formed in older cultures on SNA,<br />

PDA <strong>and</strong> also MEA (according to McKemy & Morgan-Jones 1991),<br />

but more prominent on PDA <strong>and</strong> MEA, resembling those formed<br />

by the stem-rotting morph (i.e., the periconioid morph, in planta),<br />

consisting of a long unbranched stipe <strong>and</strong> a subglobose head, but<br />

in culture the heads are often more loosely branched than on the<br />

natural substratum, not always forming a compact head, up to 580<br />

µm long, 5–13 µm wide, attenuated towards the apex, 4–8 µm<br />

just below the branched upper part, somewhat swollen at the base,<br />

septate, medium to very dark brown, minutely verruculose, walls<br />

distinctly thickened, two distinct wall layers visible, 1–2 µm thick.<br />

Conidiogenous cells holoblastic, integrated, terminal, intercalary or<br />

even discrete, ellipsoid to cylindrical or doliiform, subdenticulate,<br />

proliferation sympodial, multilocal, conidiogenous loci truncate,<br />

flat, unthickened, 1–3 µm wide, somewhat darkened-refractive; in<br />

culture conidiogenous loci appearing to be somewhat thickened<br />

<strong>and</strong> distinctly darkened-refractive, 1–2.5(–3) µm wide. Conidia<br />

catenate, in long, branched chains, straight, subglobose, aseptate,<br />

3.5–7 µm diam, or ellipsoid-ovoid, 6–15 × 4–9 µm, 0(–1)-septate,<br />

pale olivaceous to olivaceous-brown, smooth to verruculose (under<br />

the light microscope), hila flat, truncate, unthickened, (0.5–)1–<br />

2(–2.5) µm wide, not darkened, but somewhat refractive; in vitro<br />

conidia numerous, catenate, formed in long, branched chains, small<br />

conidia globose to subglobose, (2–)3–7 × (2–)3–4 µm, aseptate,<br />

intercalary ones ellipsoid-ovoid, 6–16 × 3.5–5 µm, 0(–1)-septate,<br />

secondary ramoconidia ellipsoid to cylindrical-oblong, (13–)<br />

15–34(–47) × (3–)4–6(–7) µm, 0–2-septate, sometimes slightly<br />

constricted at the septa, medium olivaceous-brown, verruculose or<br />

irregularly rough-walled, walls slightly to distinctly thickened, hila<br />

more or less protuberant, subdenticulate to denticulate, in small <strong>and</strong><br />

intercalary conidia 0.5–1(–1.5) µm, in secondary ramoconidia 1–2.5<br />

(–3) µm, unthickened or somewhat thickened, darkened-refractive;<br />

occasional microcyclic conidiogenesis.<br />

Cultural characteristics: Colonies on PDA at first whitish or smoke<br />

grey, reverse smoke-grey to olivaceous-grey, with age smoke-grey<br />

to olivaceous or olivaceous-grey, sometimes even dark mouse-grey,<br />

reverse iron-grey to dark mouse-grey or black, felty; margin white<br />

to smoke-grey, narrow to more or less broad, regular to slightly<br />

undulate, glabrous to somewhat feathery; aerial mycelium at first<br />

mainly in the colony centre, with age abundantly formed, covering<br />

almost the whole colony, whitish, smoke-grey to olivaceous, felty;<br />

growth low convex to raised; numerous small exudates formed,<br />

sometimes becoming prominent; fertile.<br />

Specimens examined: Czechoslovakia, Bohemia, Turnau, on leaves of Paeonia<br />

arborea, 15 Sep. 1905, J.E. Kabát, Kabát & Bubák, Fungi Imperf. Exs. 396, B 70-<br />

6669. France, on dead stems of Paeonia sp., 1901, ex Herbario Musei Parisiensis,<br />

ex herb. Magnus, exs. Desmazières, Pl. Crypt. N. France, Ed. 2, Ser. 1, 1621, HBG,<br />

as “Periconia atra”; Chailly-en-Biere, Seine-et-Marne, Feuilleaubois, on stems of<br />

P. officinalis, 27 Mar. 1881, Roumeguère, Fungi Sel. Gall. Exs. 1803, HBG, as<br />

“Periconia atra”. Germany, Baden-Würtemberg, Kreis Tübingen, Drusslingen, on<br />

leaves of P. officinalis, Jun. 1935, Raabe, B 70-6670; Bayern, Freising, on leaves<br />

of P. officinalis, Sep. 1918, Prof. Dr. J.E. Weiß, Herbarium pathologicum, B 70-<br />

6663; Br<strong>and</strong>enburg, Schloßpark zu Tamsel, on leaves of P. officinalis, 15 Aug.<br />

1924, P. Vogel, Sydow, Mycoth. Germ. 2447, M-57751, PH; Triglitz, on leaves of<br />

P. officinalis, 3 Oct. 1909, Jaap, B 70-6668; Hessen, Frankfurt am Main, botanical<br />

garden, on leaves of P. potaninii, 7 Oct. 2004, R. Kirschner, HAL, RoKi 2222; Kreis<br />

Kassel, Hofgeismar, Garten von Prof. Grupe, on leaves of P. officinalis, 3 Sep.<br />

1947, Schulz, B 70-6658; Mecklenburg-Vorpommern, Rostock, neuer botanischer<br />

Garten, on leaves of P. corallina (= P. mascula), 27 Aug. 1950, Becker, B 70-6662;<br />

Nordrhein-Westfalen, Duisburg, Dinslake, private garden, on leaves of P. anomala,<br />

9 Aug. 2005, N. Ale-Agha, HAL 2014 F; Hamborn, botanical garden, on leaves of P.<br />

obovata, 10 Aug. 2005, N. Ale-Agha, HAL 2017 F; Essen, botanical garden of the<br />

university of Essen, on leaves of P. mlokosewitschii, 10 Aug. 2005, N. Ale-Agha,<br />

HAL 2013 F; on leaves of P. officinalis <strong>and</strong> P. suffruticosa, 11 Aug. 2005, N. Ale-<br />

Agha, HAL 2016, 2017 F; Sachsen, Königstein, in Gärten, verbreitet, on leaves of<br />

P. officinalis, Aug. 1896, W. Krieger, Krieger, Fungi Saxon. Exs. 1545, M-57749;<br />

Aug., Sep. 1896, 1915, W. Krieger, Krieger, Schädliche Pilze, B 70-6666, 70-6667;<br />

Sachsen-Anhalt, Halle (Saale), Botanical Garden, on leaves of P. delavayi, 22<br />

Jun. 2004, K. Schubert, HAL 2011 F, culture deposited at the <strong>CBS</strong>, <strong>CBS</strong> 121522 =<br />

CPC 11383; on leaves of P. officinalis, 22 Jun. 2004, K. Schubert, HAL 2012 F; on<br />

stems of P. officinalis, 16 Mar. 2005, K. Schubert, neotype of Dichocladosporium<br />

chlorocephalum designated here HAL 1924 F, isoneotype <strong>CBS</strong>-H 19869, culture<br />

ex-neotype <strong>CBS</strong> 121523 = CPC 11969; on dead stems of Paeonia sp., Jan. 1873,<br />

G. Winter, Rabenhorst, Fungi Eur. Exs. 1661, HBG, as “Periconia chlorocephala”;<br />

Thüringen, Fürstlicher Park zu Sondershausen, on leaves of P. arborea, 20 Aug.<br />

1903, G. Oertel, Sydow, Mycoth. Germ. 196, PH. Italy, Thümen, Herb. Mycol.<br />

Oecon. 416, on living leaves of Paeonia lactiflora [= P. edulis] (M-57753), lectotype<br />

of “<strong>Cladosporium</strong> paeoniae” designated here; isolectotypes: Thümen, Herb.<br />

Mycol. Oecon. 416; Padova, on leaves of P. officinalis, Aug. 1902, P.A. Saccardo,<br />

Saccardo, Mycoth. Ital. 1186, B 70-6660, SIENA; Parma, on leaves of P. officinalis,<br />

Jul. 1876, Prof. Passerini, Thümen, Mycoth. Univ. 670, B 70-6654, 70-6655, M-<br />

57752; Pavia, botanical garden, on leaves of P. officinalis, summer 1889, Briosi &<br />

Cavara, Fung. Paras. Piante Colt. Utili Ess. 78, M-57748; F. Cavara, Roumeguère,<br />

Fungi Sel. Gall. Exs. 5193, mixed infection with <strong>Cladosporium</strong> herbarum, B 70-<br />

6656; Siena, Hort. Bot., on leaves of Paeonia sp., Nov. 1899, SIENA. Latvia, prov.<br />

Vidzeme, Kreis Riga, Riga, in a garden, on leaves of P. foemina [= P. officinalis], 28<br />

Aug. 1936, J. Smarods, Fungi Lat. Exs. 799, M-57747. New Zeal<strong>and</strong>, isolated from<br />

www.studiesinmycology.org<br />

101


Schubert et al.<br />

Fig. 5. Dichocladosporium chlorocephalum (<strong>CBS</strong> 121522 = CPC 11383). A–B. Symptoms on leaves of Paeonia officinalis <strong>and</strong> P. delavayi caused by the cladosporioid, leaf spotting<br />

morph. C–D. Conidiophores <strong>and</strong> conidia. E. Ramoconidia <strong>and</strong> conidia. F–G. Scanning electron microscopic photographs. F. Conidial chain still attached to a conidiophore.<br />

G. Conidia showing surface ornamentation <strong>and</strong> scar structure. Scale bars: C–E, G = 10 µm; F = 5 µm.<br />

red leaf <strong>and</strong> stem lesions on Paeonia sp., M. Braithwaite, <strong>CBS</strong> 100405. Romania,<br />

Râmnicu-Vâlcea, distr. Vâlcea, Oltenia, on leaves of P. officinalis, 17 Aug. 1930,<br />

Tr. Săvulescu & C. S<strong>and</strong>u, Săvulescu, Herb. Mycol. Roman. 298, M-57742. U.K.,<br />

Engl<strong>and</strong>, Cornwall, Lambounce Hill, Perranzubuloe, isolated from dead stems of<br />

Paeonia sp., isol. F. Rilstone, <strong>CBS</strong> 213.73 = IMI 048108a. U.S.A., on leaves of<br />

Paeonia sp., Sep. 1878, Ellis, N. Amer. Fungi 543, B 70-6659, M-57744, PH; Illinois,<br />

Cobden, on leaves of Paeonia sp., 8 Aug. 1882, F.S. Earle, No. 91, B 70-6657;<br />

Kansas, Topeka, on leaves of P. officinalis, 7 Jul. 1922, C.F. Menninger, US Dept.<br />

Agric., Pathol. Mycol. Coll. 60085, B 70-6661, F; Montana, Columbia, on leaves of<br />

P. officinalis, Aug. 1886, B.T. Galloway, Ellis & Everh., N. Amer. Fungi Ser. II, 1991,<br />

PH; on leaves of Paeonia sp., 18 Oct. 1931, W.E. Maneval, F.<br />

Host range <strong>and</strong> distribution: On Paeonia anomala, P. arborea, P.<br />

delavayi, P. hybrida, P. lactiflora, P. mascula, P. mlokosewitschii,<br />

P. moutan, P. obovata, P. obovata var. willmotiae, P. officinalis,<br />

P. potaninii, P. suffruticosa, Paeonia spp. (Paeoniaceae), Asia<br />

(Armenia, China, Georgia, Kazakhstan, Russia), Europe (Belgium,<br />

102


Dichocladosporium gen. nov.<br />

Czechoslovakia, Denmark, France, Germany, Italy, Latvia, Moldova,<br />

Pol<strong>and</strong>, Romania, Switzerl<strong>and</strong>, U.K., Ukraine), North America<br />

(Canada, U.S.A.), New Zeal<strong>and</strong>.<br />

Notes: Type material of Periconia chlorocephala is not preserved<br />

in the herbarium of G. Fresenius at FR (Forschungsinstitut<br />

Senkenberg, Frankfurt a. M., Germany). Hence, a new specimen<br />

collected in the Botanical Garden of the Martin-Luther-University<br />

Halle (Saale), Germany, is proposed to serve as neotype. A culture<br />

derived from this collection is deposited at the <strong>CBS</strong>, Utrecht, the<br />

Netherl<strong>and</strong>s as ex-neotype culture. A leaf-blotch sample, also<br />

collected in the Botanical Garden at Halle (Saale), from which we<br />

also derived a living culture, is designated as representative of the<br />

synanamorph, <strong>Cladosporium</strong> paeoniae. Both cultures have been<br />

used to generate DNA sequence data.<br />

<strong>The</strong> two stages (morphs) of this fungus are usually ecologically<br />

<strong>and</strong> seasonally separated, but sometimes conidiophores of the<br />

leaf-blotching (cladosporioid) morph also occur on dead stems of<br />

peony intermixed with the macronematous conidiophores of the<br />

periconioid morph. In culture conidiophore <strong>and</strong> conidial width tends<br />

to be narrower than on the natural substratum, <strong>and</strong> the conidia are<br />

not as frequently septate.<br />

DISCUSSION<br />

Cultural studies by ourselves <strong>and</strong> McKemy & Morgan-Jones<br />

(1991), <strong>and</strong> molecular sequence analyses documented herein<br />

clearly demonstrate that <strong>Cladosporium</strong> chlorocephalum, occurring<br />

on necrotic stems, <strong>and</strong> C. paeoniae, causing leaf-blotch symptoms<br />

on living leaves of Paeonia spp., are two synanamorphs of a single<br />

species, which has to be excluded from <strong>Cladosporium</strong> s. str. since<br />

the conidiogenous loci are quite distinct from the characteristically<br />

coronate scars in the latter <strong>genus</strong> <strong>and</strong> because ITS sequences<br />

indicate clear separation from <strong>Cladosporium</strong> s. str.<br />

Analysis of the morphology <strong>and</strong> conidiogenesis showed that<br />

the macronematous stage of this fungus (C. chlorocephalum, the<br />

periconioid morph) closely resembles Metulocladosporiella, recently<br />

introduced for the <strong>Cladosporium</strong> speckle disease of banana. <strong>The</strong>re<br />

are, however, some differences. In Metulocladosporiella musae<br />

(E.W. Mason) Crous et al., the type species, micronematous<br />

conidiophores occur in vitro <strong>and</strong> in vivo, <strong>and</strong> macronematous<br />

conidiophores occur on leaf-spots, whereas in C. chlorocephalum<br />

the semi-macronematous conidiophores usually accompany<br />

leaf-blotch symptoms on living leaves <strong>and</strong> the macronematous<br />

conidiophores occur in saprobic growth on old necrotic stems.<br />

Rhizoid hyphae arising from the swollen basal cells of the<br />

macronematous conidiophores are characteristic for M. musae, but<br />

lacking in C. chlorocephalum, <strong>and</strong> the conidia in the latter species<br />

are 0–5-septate, but only 0(–1)-septate in M. musae. <strong>The</strong> semimacronematous,<br />

leaf-blotching stage (the cladosporioid morph)<br />

is barely distinguishable from the present concept of Fusicladium,<br />

which includes species with catenate conidia (Schubert et al. 2003).<br />

However, the peony fungus does not cluster within the Venturiaceae.<br />

Since C. chlorocephalum clusters apart of the Chaetothyriales,<br />

the clade to which Metulocladosporiella belongs, the differences<br />

observed here seem to be sufficient to place this fungus in a new<br />

<strong>genus</strong> (also see Crous et al. 2007 – this volume). Crous et al.<br />

(2006a) discussed differences between Metulocladosporiella <strong>and</strong><br />

allied <strong>dematiaceous</strong> hyphomycete genera <strong>and</strong> provided a key to<br />

the latter <strong>genus</strong> <strong>and</strong> morphologically <strong>similar</strong> genera. Using this key,<br />

attempts to determine the macronematous morph of <strong>Cladosporium</strong><br />

www.studiesinmycology.org<br />

chlorocephalum lead to Metulocladosporiella. Differences between<br />

morphologically <strong>similar</strong> genera have been discussed in the paper<br />

by Crous et al. (2006a) <strong>and</strong> are also valid for the new <strong>genus</strong><br />

Dichocladosporium. Parapericoniella U. Braun, Heuchert & K.<br />

Schub., a fungicolous <strong>genus</strong> recently introduced to accommodate<br />

<strong>Cladosporium</strong> asterinae Deighton, is also morphologically <strong>similar</strong><br />

in having apically, densely branched conidiophores <strong>and</strong> truncate,<br />

unthickened conidiogenous loci <strong>and</strong> hila, but is quite distinct in not<br />

having micronematous conidiophores (Heuchert et al. 2005).<br />

ACKNOWLEDGEMENTS<br />

We are much obliged to the curators of B, F, HBG, M, PH <strong>and</strong> SIENA for the<br />

loans of the collections studied. R. Kirschner <strong>and</strong> N. Ale-Agha are thanked for<br />

sending collections <strong>and</strong> cultures on Paeonia spp. We are very grateful to the<br />

Institute of Zoology of the Martin-Luther-University, above all to G. Tschuch, for<br />

providing access to SEM facilities. This work was supported in part by a grant of<br />

the “Graduiertenförderung des L<strong>and</strong>es Sachsen-Anhalt” <strong>and</strong> a grant of Synthesys<br />

(No. 2559) awarded to K.S. We thank M. Vermaas for preparing the photographic<br />

plates.<br />

REFERENCES<br />

Braun U, Crous PW, Dugan FM, Groenewald JZ, Hoog GS de (2003). Phylogeny<br />

<strong>and</strong> taxonomy of cladosporium-like hyphomycetes, including Davidiella gen.<br />

nov., the teleomorph of <strong>Cladosporium</strong> s.str. Mycological Progress 2: 3–18.<br />

Crous PW (1998). Mycosphaerella spp. <strong>and</strong> their anamorphs associated with leaf<br />

spot diseases of Eucalyptus. Mycologia Memoir 21: 1–170.<br />

Crous PW, Braun U, Schubert K, Groenewald JZ (2007). Delimiting <strong>Cladosporium</strong><br />

from morphologically <strong>similar</strong> genera. Studies in Mycology 58: 33–56.<br />

Crous PW, Schroers HJ, Groenewald JZ, Braun U, Schubert K (2006a).<br />

Metulocladosporiella gen. nov. for the causal organism of <strong>Cladosporium</strong><br />

speckle disease of banana. Mycological Research 110: 264–275.<br />

Crous PW, Slippers B, Wingfield MJ, Rheeder J, Marasas WFO, Phillips AJL, Alves<br />

A, Burgess T, Barber P, Groenewald JZ (2006b). Phylogenetic lineages in the<br />

Botryosphaeriaceae. Studies in Mycology 55: 235–253.<br />

David JC (1997). A contribution to the systematics of <strong>Cladosporium</strong>. Revision of the<br />

fungi previously referred to Heterosporium. Mycological Papers 172: 1–157.<br />

Dhingra OD, Sinclair JB (1985). Basic plant pathology methods. CRC Press, Boca<br />

Raton, Florida.<br />

Fresenius JBGW (1850). Beiträge zur Mykologie 1. Heinrich Ludwig Brömmer<br />

Verlag, Frankfurt.<br />

Gams W, Verkley GJM, Crous PW (2007). <strong>CBS</strong> Course of Mycology. 5 th ed. <strong>CBS</strong>,<br />

Utrecht.<br />

Heuchert B, Braun U, Schubert K (2005). Morphotaxonomic revision of fungicolous<br />

<strong>Cladosporium</strong> species (hyphomycetes). Schlechtendalia 13: 1–78.<br />

Hoog GS de, Gerrits van den Ende AHG (1998). Molecular diagnostics of clinical<br />

strains of filamentous Basidiomycetes. Mycoses 41: 183–189.<br />

Kiffer E, Morelet M (1999). <strong>The</strong> Deuteromycetes. Mitosporic Fungi, Classification<br />

<strong>and</strong> Generic Key. Science Publishers, Enfield, NJ.<br />

Lee SB, Taylor JW (1990). Isolation of DNA from fungal mycelia <strong>and</strong> single spores.<br />

In: PCR Protocols: a guide to methods <strong>and</strong> applications (Innis MA, Gelf<strong>and</strong> DH,<br />

Sninsky JJ, White TJ, eds). Academic Press, San Diego, California: 282–287.<br />

Mason EW, Ellis MB (1953). British species of Periconia. Mycological Papers 56:<br />

1–127.<br />

McKemy JM, Morgan-Jones G (1991). Studies in the <strong>genus</strong> <strong>Cladosporium</strong> sensu lato<br />

III. Concerning <strong>Cladosporium</strong> chlorocephalum <strong>and</strong> its synonym <strong>Cladosporium</strong><br />

paeoniae, the causal organism of leaf-blotch of peony. Mycotaxon 41: 135–<br />

146.<br />

Meuli LJ (1937). <strong>Cladosporium</strong> leaf blotch of peony. Phytopathology 27: 172–182.<br />

Moncalvo J-M, Rehner SA, Vilgalys R (1993). Systematics of Lyophyllum section<br />

Difformia based on evidence from culture studies <strong>and</strong> ribosomal DNA<br />

sequences. Mycologia 85: 788–794.<br />

Passerini G (1876). <strong>Cladosporium</strong> paeoniae. Just’s Botanische Jahresberichte 4:<br />

235.<br />

Rayner RW (1970). A mycological colour chart. CMI <strong>and</strong> British Mycological Society.<br />

Kew.<br />

Rehner SA, Samuels GJ (1994). Taxonomy <strong>and</strong> phylogeny of Gliocladium analysed<br />

from nuclear large subunit ribosomal DNA sequences. Mycological Research<br />

98: 625–634.<br />

103


Schubert et al.<br />

Saccardo PA (1882). Fungorum extra-europaeorum Pugillus. Michelia 2(6): 136–<br />

149 ‘1880’.<br />

Saccardo PA (1886). Sylloge Fungorum vol. 4. Padova.<br />

Schubert K (2005). Morphotaxonomic revision of foliicolous <strong>Cladosporium</strong> species<br />

(hyphomycetes). Ph.D. dissertation. Martin-Luther-University Halle-Wittenberg.<br />

http://sundoc.bibliothek.uni-halle.de/diss-online/05/05H208/index.htm<br />

Schubert K, Ritschel A, Braun U (2003). A monograph of Fusicladium s. lat.<br />

(hyphomycetes). Schlechtendalia 9: 1–132.<br />

Vilgalys R, Hester M (1990). Rapid genetic identification <strong>and</strong> mapping of<br />

enzymatically amplified ribosomal DNA from several Cryptococcus species.<br />

Journal of Bacteriology 172: 4238–4246.<br />

Vries GA de (1952). Contribution to the knowledge of the <strong>genus</strong> <strong>Cladosporium</strong> Link<br />

ex Fr. <strong>CBS</strong>, Baarn.<br />

White TJ, Bruns T, Lee S, Taylor J (1990). Amplification <strong>and</strong> direct sequencing of<br />

fungal ribosomal RNA genes for phylogenetics. In: PCR Protocols: a guide to<br />

methods <strong>and</strong> applications (Innis MA, Gelf<strong>and</strong> DH, Sninsky JJ, White TJ, eds).<br />

Academic Press, San Diego, California: 315–322.<br />

104


available online at www.studiesinmycology.org<br />

doi:10.3114/sim.2007.58.05<br />

Studies in Mycology 58: 105–156. 2007.<br />

Biodiversity in the <strong>Cladosporium</strong> herbarum complex (Davidiellaceae, Capnodiales),<br />

with st<strong>and</strong>ardisation of methods for <strong>Cladosporium</strong> taxonomy <strong>and</strong> diagnostics<br />

K. Schubert 1* , J. Z. Groenewald 2 , U. Braun 3 , J. Dijksterhuis 2 , M. Starink 2 , C.F. Hill 4 , P. Zalar 5 , G.S. de Hoog 2 <strong>and</strong> P.W. Crous 2<br />

1<br />

Botanische Staatssammlung München, Menzinger Strasse 67, D-80638 München, Germany; 2 <strong>CBS</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD Utrecht, <strong>The</strong> Netherl<strong>and</strong>s;<br />

3<br />

Martin-Luther-Universität, Institut für Biologie, Geobotanik und Botanischer Garten, Herbarium, Neuwerk 21, D-06099 Halle (Saale), Germany; 4 Plant & Environment Laboratory<br />

Biosecurity NZ, Ministry of Agriculture & Forestry, P.O. Box 2095, Auckl<strong>and</strong> 1140, New Zeal<strong>and</strong>; 5 Biotechnical Faculty, Department of Biology, Večna pot 111, SI-1000 Ljubljana,<br />

Slovenia<br />

*Correspondence: Konstanze Schubert, konstanze.schubert@gmx.de<br />

Abstract: <strong>The</strong> <strong>Cladosporium</strong> herbarum complex comprises five species for which Davidiella teleomorphs are known. <strong>Cladosporium</strong> herbarum s. str. (D. tassiana), C.<br />

macrocarpum (D. macrocarpa) <strong>and</strong> C. bruhnei (D. allicina) are distinguishable by having conidia of different width, <strong>and</strong> by teleomorph characters. Davidiella variabile is<br />

introduced as teleomorph of C. variabile, a homothallic species occurring on Spinacia, <strong>and</strong> D. macrospora is known to be the teleomorph of C. iridis on Iris spp. <strong>The</strong> C. herbarum<br />

complex combines low molecular distance with a high degree of clonal or inbreeding diversity. Entities differ from each other by multilocus sequence data <strong>and</strong> by phenetic<br />

differences, <strong>and</strong> thus can be interpreted to represent individual taxa. Isolates of the C. herbarum complex that were formerly associated with opportunistic human infections,<br />

cluster with C. bruhnei. Several species are newly described from hypersaline water, namely C. ramotenellum, C. tenellum, C. subinflatum, <strong>and</strong> C. herbaroides. <strong>Cladosporium</strong><br />

pseudiridis collected from Iris sp. in New Zeal<strong>and</strong>, is also a member of this species complex <strong>and</strong> shown to be distinct from C. iridis that occurs on this host elsewhere in the<br />

world. A further new species from New Zeal<strong>and</strong> is C. sinuosum on Fuchsia excorticata. <strong>Cladosporium</strong> antarcticum is newly described from a lichen, Caloplaca regalis, collected<br />

in Antarctica, <strong>and</strong> C. subtilissimum from grape berries in the U.S.A., while the new combination C. ossifragi, the oldest valid name of the <strong>Cladosporium</strong> known from Narthecium<br />

in Europe, is proposed. St<strong>and</strong>ard protocols <strong>and</strong> media are herewith proposed to facilitate future morphological examination of <strong>Cladosporium</strong> spp. in culture, <strong>and</strong> neotypes or<br />

epitypes are proposed for all species treated.<br />

Taxonomic novelties: <strong>Cladosporium</strong> antarcticum K. Schub., Crous & U. Braun, sp. nov., C. herbaroides K. Schub., Zalar, Crous & U. Braun, sp. nov., C. ossifragi (Rostr.) U.<br />

Braun & K. Schub., comb. nov., C. pseudiridis K. Schub., C.F. Hill, Crous & U. Braun, sp. nov., C. ramotenellum K. Schub., Zalar, Crous & U. Braun, sp. nov., C. sinuosum K.<br />

Schub., C.F. Hill, Crous & U. Braun, sp. nov., C. subinflatum K. Schub., Zalar, Crous & U. Braun, sp. nov., C. subtilissimum K. Schub., Dugan, Crous & U. Braun, sp. nov., C.<br />

tenellum K. Schub., Zalar, Crous & U. Braun sp. nov., Davidiella macrocarpa Crous, K. Schub. & U. Braun, sp. nov., D. variabile Crous, K. Schub. & U. Braun, sp. nov.<br />

Key words: Clonality, Davidiella, homothallism, new species, phylogeny, recombination, taxonomy.<br />

Introduction<br />

<strong>Cladosporium</strong> herbarum (Pers. : Fr.) Link, type species of the <strong>genus</strong><br />

<strong>Cladosporium</strong> Link, is one of the most common environmental fungi<br />

to be isolated worldwide. It abundantly occurs on fading or dead<br />

leaves of herbaceous <strong>and</strong> woody plants, as secondary invader<br />

on necrotic leaf spots, <strong>and</strong> has frequently been isolated from<br />

air (Samson et al. 2000), soil (Domsch et al. 1980), foodstuffs,<br />

paints, textiles, humans (de Hoog et al. 2000) <strong>and</strong> numerous<br />

other substrates. It is also known to occur on old carpophores<br />

of mushrooms <strong>and</strong> other fungi (Heuchert et al. 2005) <strong>and</strong> to be<br />

a common endophyte (Riesen & Sieber 1985, Brown et al. 1998,<br />

El-Morsy 2000), especially in temperate regions. Under favourable<br />

climatic conditions C. herbarum also germinates <strong>and</strong> grows as an<br />

epiphyte on the surface of green, healthy leaves (Schubert 2005).<br />

Persoon (1794) introduced C. herbarum as Dematium<br />

herbarum Pers., which was later reclassified by Link (1809) as<br />

Acladium herbarum (Pers.) Link. In 1816, Link included C. herbarum<br />

together with three additional species in his newly described <strong>genus</strong><br />

<strong>Cladosporium</strong>. Clements & Shear (1931) proposed C. herbarum<br />

as lectotype species of the latter <strong>genus</strong>, a decision followed by de<br />

Vries (1952) <strong>and</strong> Hughes (1958). Several authors provided detailed<br />

treatments of C. herbarum (de Vries 1952, Ellis 1971, Domsch et al.<br />

1980, Prasil & de Hoog 1988), <strong>and</strong> there are literally thous<strong>and</strong>s of<br />

records of this species in the literature. McKemy & Morgan-Jones<br />

(1991) <strong>and</strong> Ho et al. (1999) examined C. herbarum in culture <strong>and</strong><br />

published detailed descriptions of its features in vitro.<br />

<strong>Cladosporium</strong> macrocarpum Preuss, a second component<br />

within the herbarum complex, has hitherto been known <strong>and</strong> treated<br />

as an allied, but morphologically distinct species on the basis of its<br />

wider <strong>and</strong> somewhat larger, frequently 2–3-septate, more regularly<br />

verrucose conidia, shorter conidial chains <strong>and</strong> more pronounced<br />

prolongations of the conidiophores. Dugan & Roberts (1994) carried<br />

out examinations of morphological <strong>and</strong> reproductive aspects of both<br />

species, <strong>and</strong> in so doing demonstrated a morphological continuum<br />

between C. macrocarpum <strong>and</strong> C. herbarum, concluding that the<br />

name herbarum should have preference. <strong>The</strong>refore, Ho et al. (1999)<br />

introduced the new combination C. herbarum var. macrocarpum<br />

(Preuss) M.H.M. Ho & Dugan. Although transitional forms have<br />

been discussed to occur between the two species, several authors<br />

still prefer to retain C. macrocarpum as a separate species.<br />

In an attempt to elucidate the species within the C. herbarum<br />

complex, therefore, a multilocus DNA sequence typing approach<br />

was used, employing five genes, namely the internal transcribed<br />

spacers of the rDNA genes (ITS), actin, calmodulin, translation<br />

elongation factor 1-α, <strong>and</strong> histone H3. <strong>The</strong>se data were<br />

supplemented with morphological examinations under st<strong>and</strong>ardised<br />

conditions, using light <strong>and</strong> scanning electron microscopy, as well as<br />

cultural characteristics <strong>and</strong> growth studies.<br />

Material <strong>and</strong> methods<br />

Isolates<br />

Isolates included in this study were obtained from the culture<br />

collection of the Centraalbureau voor Schimmelcultures (<strong>CBS</strong>),<br />

Utrecht, Netherl<strong>and</strong>s, or were freshly isolated from a range of<br />

different substrates. Single-conidial <strong>and</strong> ascospore isolates were<br />

obtained using the techniques as explained in Crous (1998) for<br />

species of Mycosphaerella Johanson <strong>and</strong> its anamorphs. Isolates<br />

105


Schubert et al.<br />

were inoculated onto 2 % potato-dextrose agar (PDA), synthetic<br />

nutrient-poor agar (SNA), 2 % malt extract agar (MEA) <strong>and</strong> oatmeal<br />

agar (OA) (Gams et al. 2007), <strong>and</strong> incubated under continuous<br />

near-ultraviolet light at 25 °C to promote sporulation. All cultures<br />

obtained in this study are maintained in the culture collection of<br />

the <strong>CBS</strong> (Table 1). Nomenclatural novelties <strong>and</strong> descriptions were<br />

deposited in MycoBank (www.MycoBank.org).<br />

DNA isolation, amplification <strong>and</strong> sequence analysis<br />

Fungal colonies were established on agar plates, <strong>and</strong> genomic<br />

DNA was isolated as described in Gams et al. (2007). Partial gene<br />

sequences were determined as described by Crous et al. (2006)<br />

for actin (ACT), calmodulin (CAL), translation elongation factor 1-<br />

alpha (EF), histone H3 (HIS) <strong>and</strong> part (ITS) of the nuclear rDNA<br />

operon spanning the 3’ end of the 18S rRNA gene (SSU), the<br />

first internal transcribed spacer, the 5.8S rRNA gene, the second<br />

internal transcribed spacer <strong>and</strong> the 5’ end of the 28S rRNA gene<br />

(LSU). <strong>The</strong> nucleotide sequences were generated using both PCR<br />

primers to ensure good quality sequences over the entire length of<br />

the amplicon. Subsequent sequence alignment <strong>and</strong> phylogenetic<br />

analysis followed the methods of Crous et al. (2006). Gaps longer<br />

than 10 bases were coded as single events for the phylogenetic<br />

analyses; the remaining gaps were treated as new character<br />

states. Sequence data were deposited in GenBank (Table 1) <strong>and</strong><br />

the alignment <strong>and</strong> tree in TreeBASE (www.treebase.org).<br />

Data analysis<br />

<strong>The</strong> number of entities in the dataset of 79 strains was inferred<br />

with Structure v. 2.2 software (Pritchard et al. 2000, Falush et al.<br />

2003) using an UPGMA tree of data of the ACT gene compared<br />

with CAL, EF <strong>and</strong> HIS with the exclusion of the nearly invariant ITS<br />

region. For this analysis group indications were derived from a tree<br />

produced with MrAic (Nyl<strong>and</strong>er 2004). <strong>The</strong> length of the burn-in<br />

period was set to 1 000 000, number of MCMC repeats after burn-in<br />

10 000, with admixture ancestry <strong>and</strong> allele frequencies correlated<br />

models, assuming that all groups diverged from a recent ancestral<br />

population <strong>and</strong> that allele frequencies are due to drift. Uniform prior<br />

for ALPHA was set to 1.0 (default) <strong>and</strong> allele frequencies with λ set to<br />

1.0 (default). <strong>The</strong> numbers of MCMC repetitions after burn-in were<br />

set as 10 000 <strong>and</strong> 100 000. <strong>The</strong> number of clusters (K) in Structure<br />

was assumed from 5 to 7. Population differentiation F ST<br />

(index: θ)<br />

was calculated with 1–6 runs using the same software. <strong>The</strong> null<br />

hypothesis for this analysis is no population differentiation. When<br />

observed theta (θ) is significantly different from those of r<strong>and</strong>om<br />

data sets (p < 0.05), population differentiation is considered.<br />

Association of multilocus genotypes was screened with the<br />

multilocus option in BioNumerics v. 4.5. To test for reproductive<br />

mode in each population, the st<strong>and</strong>ardised index of association<br />

(I S ; Haubold et al. 1998) was calculated with start2 software<br />

A<br />

(Jolley et al. 2001). <strong>The</strong> null hypothesis for this analysis is complete<br />

panmixia. <strong>The</strong> values of I S were compared between observed <strong>and</strong><br />

A<br />

r<strong>and</strong>omised datasets. <strong>The</strong> hypothesis would be rejected when p <<br />

0.05. Mean genetic diversity (H) <strong>and</strong> diversities of individual loci<br />

were calculated with lian v. 3.5 (Haubold & Hudson 2000). Degrees<br />

of recombination or horizontal gene transfer were also visualised<br />

using SplitsTree v. 4.8 software (Huson & Bryant 2006). Split<br />

decomposition was carried out with default settings, i.e., character<br />

transformation using uncorrected (observed, “P”) distances, splits<br />

transformation using “equal angle”, <strong>and</strong> optimise boxes iteration set<br />

to 2.<br />

Morphology<br />

As the present study represents the first in a series dealing<br />

with <strong>Cladosporium</strong> spp. <strong>and</strong> their Davidiella Crous & U. Braun<br />

teleomorphs in culture, a specific, st<strong>and</strong>ardised protocol was<br />

established by which all species complexes will be treated in<br />

future.<br />

Morphology of the anamorph: Microscopic observations were made<br />

from colonies cultivated for 7 d under continuous near-ultraviolet<br />

light at 25 °C on SNA. Preparations were mounted in Shear’s<br />

solution (Gams et al. 2007). To study conidial development <strong>and</strong><br />

branching patterns, squares of transparent adhesive tape (Titan<br />

Ultra Clear Tape, Conglom Inc., Toronto, Canada) were placed on<br />

conidiophores growing in the zone between the colony margin <strong>and</strong><br />

2 cm inwards, <strong>and</strong> mounted between two drops of Shear’s solution<br />

under a glass coverslip. Different types of conidia are formed by<br />

<strong>Cladosporium</strong> species for which different terms need to be adopted.<br />

Ramoconidia are conidia with usually more than one (mostly 2<br />

or 3) conidial hilum, which typically accumulate at the tip of these<br />

conidia. Conidiogenous cells with more than one conidiogenous<br />

locus are first formed as apical parts of conidiophores. Such apical<br />

Fig. 1. <strong>Cladosporium</strong> conidiophore with ramoconidia, secondary ramoconidia,<br />

intercalary conidia, <strong>and</strong> small, terminal conidia. Scale bar = 10 µm. K. Schubert<br />

del.<br />

106


<strong>Cladosporium</strong> herbarum species complex<br />

Aculeate<br />

Spinulose<br />

Digitate<br />

Muricate<br />

Granulate<br />

Colliculate<br />

Pustulate<br />

Pedicellate<br />

Fig. 2. Terms used to describe conidium wall ornamentation under the cryo-electron<br />

microscope. Adapted from David (1997).<br />

than the small terminal conidia. In older literature true ramoconidia<br />

were often cited as “ramoconidia s. str.”, whereas secondary<br />

ramoconidia have been referred to as “ramoconidia s. lat.”<br />

Morphology of the teleomorph: Teleomorphs were induced by<br />

inoculating plates of 2 % tap water agar onto which autoclaved<br />

stem pieces of Urtica dioica (European stinging nettle) were placed.<br />

Inoculated plates were incubated on the laboratory bench for 7 d,<br />

after that period they were further incubated at 10 °C in the dark for<br />

1–2 mo to stimulate teleomorph development. Wherever possible,<br />

30 measurements (× 1 000 magnification) were made of conidia<br />

<strong>and</strong> ascospores, with the extremes of spore measurements given<br />

in parentheses. Cultural characteristics: Colonies were cultivated<br />

on PDA, MEA <strong>and</strong> OA plates for 14 d at 25 °C in the dark, after<br />

which the surface <strong>and</strong> reverse colours were rated using the charts<br />

of Rayner (1970). Linear growth was determined on MEA, PDA <strong>and</strong><br />

OA plates by inoculating three plates per isolate for each medium,<br />

<strong>and</strong> incubating them for 14 d at 25 ºC, after that period colony<br />

diameters were determined.<br />

Low-temperature scanning electron microscopy<br />

Isolates of <strong>Cladosporium</strong> spp. were grown on SNA with 30 g agar/L<br />

for 3–4 d at room temperature under black light. Relevant parts of<br />

the small colonies with conidiophores <strong>and</strong> conidia were selected<br />

under a binocular, excised with a surgical blade as small agar (3<br />

× 3 mm) blocks, <strong>and</strong> transferred to a copper cup for snap-freezing<br />

in nitrogen slush. Agar blocks were glued to the copper surface<br />

with frozen tissue medium (KP-Cryoblock, Klinipath, Duiven,<br />

Netherl<strong>and</strong>s) mixed with 1 part colloidal graphite (Agar Scientific,<br />

Stansted, U.K.). Samples were examined in a JEOL 5600LV<br />

scanning electron microscope (JEOL, Tokyo, Japan) equipped<br />

with an Oxford CT1500 Cryostation for cryo-electron microscopy<br />

(cryoSEM). Electron micrographs were acquired from uncoated<br />

frozen samples, or after sputter-coating by means of a gold/<br />

palladium target for 3 times during 30 s (Fig. 2). Micrographs of<br />

uncoated samples were taken at an acceleration voltage of 3 kV,<br />

<strong>and</strong> consisted out of 30 averaged fast scans (SCAN 2 mode), <strong>and</strong><br />

at 5 kV in case of the coated samples (PHOTO mode).<br />

parts of conidiophores are called ramoconidia if they secede at<br />

a septum from the conidiophore (Kirk et al. 2001). <strong>The</strong> septum at<br />

which the ramoconidium secedes often appears to be somewhat<br />

refractive or darkened. Ramoconidia are characterised by having<br />

a truncate, undifferentiated base (thus they lack a differentiated,<br />

coronate basal hilum formed in the context of conidiogenesis)<br />

<strong>and</strong> they can be very long, aseptate to sometimes multi-septate.<br />

Although they were formed initially as part of the conidiophore,<br />

they function as propagules. Only few of the species known<br />

until now have the ability to form true ramoconidia. Secondary<br />

ramoconidia also have more than one distal conidial hilum but<br />

they always derive from a conidiogenous locus of an earlier formed<br />

cell, which can be either a conidiogenous cell or a ramoconidium.<br />

Secondary ramoconidia are often shorter but somewhat wider than<br />

ramoconidia; they are often septate, <strong>and</strong> typically have a narrowed<br />

base with a coronate hilum (Fig. 1). Conidia in <strong>Cladosporium</strong> are<br />

cells with a coronate basal hilum, which is formed in the context of<br />

conidiogenesis <strong>and</strong> with either a single (when formed as intercalary<br />

units in unbranched parts of chains) or without any distal conidial<br />

hilum (when formed at the tip of conidial chains). For the first, the<br />

term “intercalary conidium” <strong>and</strong> for the latter, “small terminal<br />

conidium” is used. Intercalary conidia typically are larger <strong>and</strong> more<br />

pigmented <strong>and</strong> have a more differentiated surface ornamentation<br />

RESULTS<br />

Phylogeny <strong>and</strong> differentiation<br />

<strong>The</strong> manually adjusted alignment contained 80 sequences (including<br />

the outgroup sequence) <strong>and</strong> the five loci were represented by a<br />

total of 1 516 characters including alignment gaps which were<br />

used in the analysis. Of the 1 516 characters, 369 were parsimonyinformative,<br />

259 were variable <strong>and</strong> parsimony-uninformative, <strong>and</strong><br />

888 were constant.<br />

Forty equally most parsimonious trees (TL = 1 933 steps; CI<br />

= 0.569; RI = 0.786; RC = 0.447), one of which is shown in Fig.<br />

3, were obtained from the parsimony analysis of the combined<br />

genes. Neighbour-joining analysis using three substitution models<br />

(uncorrected “p”, Kimura 2-parameter <strong>and</strong> HKY85) on the sequence<br />

data yielded trees with identical topologies. <strong>The</strong>se differed from<br />

the tree presented in Fig. 3 with regard to the placement of C.<br />

macrocarpum strain CPC 12054 which was placed as a sister<br />

branch to the C. bruhnei Linder clade in the distance analyses<br />

(results not shown) because it shares an identical CAL sequence.<br />

All cryptic species consisting of multiple strains are clustering in<br />

well-supported clades with bootstrap support values ranging from<br />

71 % (C. herbarum) to 100 % [e.g. C. ramotenellum K. Schub.,<br />

www.studiesinmycology.org<br />

107


Schubert et al.<br />

Table 1. Isolates subjected to DNA sequence analyses <strong>and</strong> morphological examinations.<br />

Anamorph Teleomorph Accession number 1 Host Country Collector GenBank numbers 2<br />

(ITS, EF, ACT, CAL, HIS)<br />

<strong>Cladosporium</strong> antarcticum — <strong>CBS</strong> 690.92* (ex-type) Caloplaca regalis Antarctica C. Möller EF679334, EF679405, EF679484, EF679560, EF679636<br />

<strong>Cladosporium</strong> bruhnei Davidiella allicina <strong>CBS</strong> 134.31 = ATCC 11283 — Germany — EF679335, EF679406, EF679485, EF679561, EF679637<br />

<strong>CBS</strong> 157.82 Quercus robur Belgium — EF679336, EF679407, EF679486, EF679562, EF679638<br />

<strong>CBS</strong> 159.54 = ATCC 36948 Man, skin <strong>The</strong> Netherl<strong>and</strong>s — EF679337, EF679408, EF679487, EF679563, EF679639<br />

<strong>CBS</strong> 161.55 Man, sputum <strong>The</strong> Netherl<strong>and</strong>s — EF679338, EF679409, EF679488, EF679564, EF679640<br />

<strong>CBS</strong> 177.71 Thuja tincture <strong>The</strong> Netherl<strong>and</strong>s — EF679339, EF679410, EF679489, EF679565, EF679641<br />

<strong>CBS</strong> 188.54 = ATCC 11290 = IMI 049638 = CPC — — — AY251077, EF679411, EF679490, EF679566, EF679642<br />

3686<br />

<strong>CBS</strong> 366.80 Man, skin <strong>The</strong> Netherl<strong>and</strong>s — EF679340, EF679412, EF679491, EF679567, EF679643<br />

<strong>CBS</strong> 399.80 Man, skin <strong>The</strong> Netherl<strong>and</strong>s — AJ244227, EF679413, EF679492, EF679568, EF679644<br />

<strong>CBS</strong> 521.68 Air <strong>The</strong> Netherl<strong>and</strong>s — EF679341, EF679414, EF679493, EF679569, EF679645<br />

<strong>CBS</strong> 572.78 Polyporus radiatus Russia V.K. Melnik DQ289799, EF679415, DQ289866, DQ289831, EF679646<br />

<strong>CBS</strong> 813.71 Polygonatum odoratum Czech Republic — EF679342, EF679416, EF679494, EF679570, EF679647<br />

<strong>CBS</strong> 110024 Industrial water Germany, Nordrhein-Westfalen — EF679343, EF679417, EF679495, EF679571, EF679648<br />

<strong>CBS</strong> 115683 = ATCC 66670 = CPC 5101 CCA-treated Douglas-fire pole U.S.A., New York C.J. Wang AY361959, EF679418, AY752193, AY752224, AY752255<br />

<strong>CBS</strong> 121624* = CPC 12211 (neotype) Hordeum vulgare Belgium J.Z. Groenewald EF679350, EF679425, EF679502, EF679578, EF679655<br />

CPC 11386 Tilia cordata Germany, Sachsen-Anhalt K. Schubert EF679344, EF679419, EF679496, EF679572, EF679649<br />

CPC 11840 Ourisia macrophylla New Zeal<strong>and</strong> A. Blouin EF679345, EF679420, EF679497, EF679573, EF679650<br />

CPC 12042 = EXF-389 Hypersaline water from salterns Slovenia P. Zalar EF679346, EF679421, EF679498, EF679574, EF679651<br />

CPC 12045 = EXF-594 Hypersaline water from salterns Spain P. Zalar EF679347, EF679422, EF679499, EF679575, EF679652<br />

CPC 12046 = EXF-680 Air conditioning system Slovenia P. Zalar EF679348, EF679423, EF679500, EF679576, EF679653<br />

CPC 12139 Hordeum vulgare <strong>The</strong> Netherl<strong>and</strong>s — EF679349, EF679424, EF679501, EF679577, EF679654<br />

CPC 12212 Hordeum vulgare Belgium J.Z. Groenewald EF679351, EF679426, EF679503, EF679579, EF679656<br />

CPC 12921 Eucalyptus sp. Australia — EF679352, EF679427, EF679504, EF679580, EF679657<br />

<strong>Cladosporium</strong> cladosporioides — <strong>CBS</strong> 673.69 Air <strong>The</strong> Netherl<strong>and</strong>s — EF679353, EF679428, EF679505, EF679581, EF679658<br />

complex<br />

Davidiella sp. <strong>CBS</strong> 109082 Silene maritima United Kingdom A. Aptroot EF679354, EF679429, EF679506, EF679582, EF679659<br />

— CPC 11606 Musa sp. India M. Arzanlou EF679355, EF679430, EF679507, EF679583, EF679660<br />

— CPC 11609 Musa sp. India M. Arzanlou EF679356, EF679431, EF679508, EF679584, EF679661<br />

<strong>Cladosporium</strong> herbaroides — <strong>CBS</strong> 121626* = CPC 12052 = EXF-1733 (ex-type) Hypersaline water from salterns Israel P. Zalar EF679357, EF679432, EF679509, EF679585, EF679662<br />

<strong>Cladosporium</strong> herbarum Davidiella tassiana <strong>CBS</strong> 111.82 Arctostaphylos uva-ursi Switzerl<strong>and</strong> E. Müller AJ238469, EF679433, EF679510, EF679586, EF679663<br />

<strong>CBS</strong> 300.49 Biscutella laevigata Switzerl<strong>and</strong> J.A. von Arx EF679358, EF679434, EF679511, EF679587, EF679664<br />

<strong>CBS</strong> 121621* = CPC 12177 (epitype) Hordeum vulgare <strong>The</strong> Netherl<strong>and</strong>s — EF679363, EF679440, EF679516, EF679592, EF679670<br />

CPC 11600 Delphinium barbeyi U.S.A., Colorado A. Ramalay DQ289800, EF679435, DQ289867, DQ289832, EF679665<br />

CPC 11601 Delphinium barbeyi U.S.A., Colorado A. Ramalay EF679359, EF679436, EF679512, EF679588, EF679666<br />

CPC 11602 Delphinium barbeyi U.S.A., Colorado A. Ramalay EF679360, EF679437, EF679513, EF679589, EF679667<br />

CPC 11603 Delphinium barbeyi U.S.A., Colorado A. Ramalay EF679361, EF679438, EF679514, EF679590, EF679668<br />

108


<strong>Cladosporium</strong> herbarum species complex<br />

<strong>Cladosporium</strong> iridis Davidiella<br />

macrospora<br />

<strong>Cladosporium</strong> macrocarpum Davidiella<br />

macrocarpa<br />

CPC 11604 Delphinium barbeyi U.S.A., Colorado A. Ramalay EF679362, EF679439, EF679515, EF679591, EF679669<br />

CPC 12178 Hordeum vulgare <strong>The</strong> Netherl<strong>and</strong>s — EF679364, EF679441, EF679517, EF679593, EF679671<br />

CPC 12179 Hordeum vulgare <strong>The</strong> Netherl<strong>and</strong>s — EF679365, EF679442, EF679518, EF679594, EF679672<br />

CPC 12180 Hordeum vulgare <strong>The</strong> Netherl<strong>and</strong>s — EF679366, EF679443, EF679519, EF679595, EF679673<br />

CPC 12181 Hordeum vulgare <strong>The</strong> Netherl<strong>and</strong>s — EF679367, EF679444, EF679520, EF679596, EF679674<br />

CPC 12183 Hordeum vulgare <strong>The</strong> Netherl<strong>and</strong>s — EF679368, EF679445, EF679521, EF679597, EF679675<br />

<strong>CBS</strong> 107.20 Iris sp. — — EF679369, EF679446, EF679522, EF679598, EF679676<br />

<strong>CBS</strong> 138.40* (epitype) Iris sp. <strong>The</strong> Netherl<strong>and</strong>s — EF679370, EF679447, EF679523, EF679599, EF679677<br />

<strong>CBS</strong> 175.82 Water Romania — EF679371, EF679448, EF679524, EF679600, EF679678<br />

<strong>CBS</strong> 223.31 = ATCC 11287 Mycosphaerella tulasnei — — AF222830, EF679449, EF679525, EF679601, EF679679<br />

<strong>CBS</strong> 299.67 Triticum aestivum Turkey — EF679372, EF679450, EF679526, EF679602, EF679680<br />

<strong>CBS</strong> 121811* = CPC 12755 (neotype) Spinacia oleracea U.S.A. — EF679376, EF679454, EF679530, EF679606, EF679684<br />

CPC 11817 Corylus sp. U.S.A. — EF679373, EF679451, EF679527, EF679603, EF679681<br />

CPC 12054 = EXF-2287 Hypersaline water from salterns Slovenia P. Zalar EF679374, EF679452, EF679528, EF679604, EF679682<br />

<strong>CBS</strong> H-19855 = CPC 12752 = <strong>CBS</strong> 121623 Spinacia oleracea U.S.A. — EF679375, EF679453, EF679529, EF679605, EF679683<br />

CPC 12756 Spinacia oleracea U.S.A. — EF679377, EF679455, EF679531, EF679607, EF679685<br />

CPC 12757 Spinacia oleracea U.S.A. — EF679378, EF679456, EF679532, EF679608, EF679686<br />

CPC 12758 Spinacia oleracea U.S.A. — EF679379, EF679457, EF679533, EF679609, EF679687<br />

CPC 12759 Spinacia oleracea U.S.A. — EF679380, EF679458, EF679534, EF679610, EF679688<br />

<strong>Cladosporium</strong> ossifragi — <strong>CBS</strong> 842.91* (epitype) Narthecium ossifragum Norway M. di Menna EF679381, EF679459, EF679535, EF679611, EF679689<br />

<strong>CBS</strong> 843.91 Narthecium ossifragum Norway M. di Menna EF679382, EF679460, EF679536, EF679612, EF679690<br />

<strong>Cladosporium</strong> pseudiridis — <strong>CBS</strong> 116463* = LYN 1065 = ICMP 15579 (ex-type) Iris sp. New Zeal<strong>and</strong> C.F. Hill EF679383, EF679461, EF679537, EF679613, EF679691<br />

<strong>Cladosporium</strong> ramotenellum — <strong>CBS</strong> 121628* = CPC 12043 = EXF-454 (ex-type) Hypersaline water from salterns Slovenia P. Zalar EF679384, EF679462, EF679538, EF679614, EF679692<br />

CPC 12047 = EXF-967 Air conditioning system Slovenia P. Zalar EF679385, EF679463, EF679539, EF679615, EF679693<br />

Fuchsia excorticata New Zeal<strong>and</strong> A. Blouin EF679386, EF679464, EF679540, EF679616, EF679694<br />

<strong>Cladosporium</strong> sinuosum — <strong>CBS</strong> 121629* = CPC 11839 = ICMP 15819 (extype)<br />

<strong>Cladosporium</strong> spinulosum — <strong>CBS</strong> 102044 Hypersaline water from salterns Slovenia S. Soujak EF679387, EF679465, EF679541, EF679617, EF679695<br />

<strong>CBS</strong> 119907* = CPC 12040 = EXF-334 (ex-type) Hypersaline water from salterns Slovenia P. Zalar EF679388, EF679466, EF679542, EF679618, EF679696<br />

<strong>Cladosporium</strong> subinflatum — <strong>CBS</strong> 121630* = CPC 12041 = EXF-343 (ex-type) Hypersaline water from salterns Slovenia P. Zalar EF679389, EF679467, EF679543, EF679619, EF679697<br />

<strong>Cladosporium</strong> sp. — <strong>CBS</strong> 172.52 = ATCC 11320 Carya illinoensis U.S.A. — EF679390, EF679468, EF679544, EF679620, EF679698<br />

<strong>CBS</strong> 113741 Grape berry U.S.A. — EF679391, EF679469, EF679545, EF679621, EF679699<br />

<strong>CBS</strong> 113742 Grape berry U.S.A. — EF679392, EF679470, EF679546, EF679622, EF679700<br />

<strong>CBS</strong> 113744 Grape bud U.S.A. — EF679393, EF679471, EF679547, EF679623, EF679701<br />

CPC 12484 Pinus ponderosa Argentina A. Greslebin EF679394, EF679472, EF679548, EF679624, EF679702<br />

CPC 12485 Pinus ponderosa Argentina A. Greslebin EF679395, EF679473, EF679549, EF679625, EF679703<br />

<strong>Cladosporium</strong> subtilissimum — <strong>CBS</strong> 113753 Bing cherry fruits U.S.A. — EF679396, EF679474, EF679550, EF679626, EF679704<br />

<strong>CBS</strong> 113754* Grape berry U.S.A. — EF679397, EF679475, EF679551, EF679627, EF679705<br />

CPC 12044 = EXF-462 Hypersaline water from salterns Slovenia P. Zalar EF679398, EF679476, EF679552, EF679628, EF679706<br />

<strong>Cladosporium</strong> tenellum — <strong>CBS</strong> 121634* = CPC 12053 = EXF-1735 (ex-type) Hypersaline water from salterns Israel P. Zalar EF679401, EF679479, EF679555, EF679631, EF679709<br />

www.studiesinmycology.org<br />

109


Schubert et al.<br />

Table 1. (Continued).<br />

Anamorph Teleomorph Accession number 1 Host Country Collector GenBank numbers 2<br />

(ITS, EF, ACT, CAL, HIS)<br />

CPC 11813 Phyllactinia sp. on Corylus sp. U.S.A. D. Glawe EF679399, EF679477, EF679553, EF679629, EF679707<br />

CPC 12051 = EXF-1083 Hypersaline water from salterns Israel P. Zalar EF679400, EF679478, EF679554, EF679630, EF679708<br />

<strong>Cladosporium</strong> variabile Davidiella variabile <strong>CBS</strong> 121636* = CPC 12751 (epitype) Spinacia oleracea U.S.A. — EF679402, EF679480, EF679556, EF679632, EF679710<br />

CPC 12753 Spinacia oleracea U.S.A. — EF679403, EF679481, EF679557, EF679633, EF679711<br />

— Davidiella sp. <strong>CBS</strong> 289.49 Allium schoenoprasum Switzerl<strong>and</strong> E. Müller AY152552, EF679482, EF679558, EF679634, EF679712<br />

<strong>CBS</strong> 290.49 Trisetum distichophyllum Switzerl<strong>and</strong> E. Müller EF679404, EF679483, EF679559, EF679635, EF679713<br />

1 ATCC: American Type Culture Collection, Virginia, U.S.A.; <strong>CBS</strong>: Centraalbureau voor Schimmelcultures, Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; CPC: Culture collection of Pedro Crous, housed at <strong>CBS</strong>; EXF: Extremophilic Fungi Culture Collection of the Department<br />

of Biology, Biotechnical Faculty, University of Ljubljana, Slovenia; ICMP: International Collection of Micro-organisms from Plants, L<strong>and</strong>care Research, Private Bag 92170, Auckl<strong>and</strong>, New Zeal<strong>and</strong>; IMI: International Mycological Institute, CABI-Bioscience,<br />

Egham, Bakeham Lane, U.K.<br />

2 ACT: partial actin gene, CAL: partial calmodulin gene, EF: partial elongation factor 1-alpha gene, HIS: partial histone H3 gene, ITS: internal transcribed spacer region.<br />

*Ex-type cultures.<br />

Zalar, Crous & U. Braun <strong>and</strong> C. ossifragi (Rostr.) U. Braun & K.<br />

Schub.]. <strong>The</strong> intraspecific variation in the C. bruhnei clade is due<br />

to genetic variation present in the sequence data of all loci except<br />

for ITS, those in the C. macrocarpum clade in all loci except for ITS<br />

<strong>and</strong> ACT, <strong>and</strong> those in the C. herbarum clade in all loci except for<br />

ITS <strong>and</strong> CAL (data not shown). However, none of the variation for<br />

these species could be linked to host specificity or morphological<br />

differences. In general, ITS data did not provide any resolution<br />

within the C. herbarum complex, whereas EF data provided species<br />

clades with very little intraspecific variation <strong>and</strong> ACT, CAL <strong>and</strong> HIS<br />

revealed increasing intraspecific variation (ACT the least <strong>and</strong> HIS<br />

the most).<br />

<strong>The</strong> mean genetic diversity (H) of the entire data set excluding<br />

the nearly invariant ITS region was 0.9307, with little difference<br />

between genes (ACT = 0.9257, CAL = 0.9289, EF = 0.9322, HIS<br />

= 0.9361). <strong>The</strong> loci showed different numbers of alleles (ACT: 22,<br />

CAL: 16, EF: 21, HIS: 20, ITS: 6). Differentiation of entities when<br />

calculated with Structure software using the admixture/correlated<br />

model showed highest value with K = 6. At this value F ST<br />

varied<br />

between 0.1362 <strong>and</strong> 0.3381. Linkage disequilibrium calculated<br />

using the st<strong>and</strong>ardised index of association (I S ) for the entire<br />

A<br />

dataset (observed variance V o<br />

= 0.5602, expected variance V e<br />

=<br />

0.2576) was 0.3914 (P = 0.0001), consistent with a small amount<br />

of recombination that did not destroy the linkage between alleles.<br />

Only few groups appeared to be separated for all alleles; degrees<br />

of gene flow are indicated in Fig. 4. SplitsTree software produced<br />

unresolved star-shaped structures for all genes, without any sign of<br />

reticulation (Fig. 5).<br />

110


<strong>Cladosporium</strong> herbarum species complex<br />

100<br />

74<br />

100<br />

10 changes<br />

Cercospora beticola CPC11557<br />

CPC 11609<br />

<strong>CBS</strong> 109082<br />

<strong>CBS</strong> 673.69<br />

<strong>Cladosporium</strong> cladosporioides complex<br />

98<br />

CPC 11606<br />

Davidiella sp. <strong>CBS</strong> 290.49<br />

100 CPC 12043<br />

CPC 12047 <strong>Cladosporium</strong> ramotenellum<br />

100 <strong>CBS</strong> 842.91<br />

80<br />

<strong>Cladosporium</strong> ossifragi<br />

<strong>CBS</strong> 843.91<br />

Davidiella sp. <strong>CBS</strong> 289.49<br />

CPC 12053<br />

91<br />

100<br />

CPC 11813 <strong>Cladosporium</strong> tenellum<br />

64<br />

CPC 12051<br />

<strong>Cladosporium</strong> sp. <strong>CBS</strong> 113744<br />

63<br />

54<br />

94<br />

100<br />

98<br />

72 62<br />

81<br />

66<br />

67<br />

54<br />

67<br />

91<br />

51<br />

<strong>CBS</strong> 113753<br />

<strong>CBS</strong> 113754<br />

CPC 12044<br />

CPC 12139<br />

<strong>CBS</strong> 159.54<br />

<strong>CBS</strong> 188.54<br />

<strong>CBS</strong> 177.71<br />

<strong>CBS</strong> 399.80<br />

<strong>CBS</strong> 521.68<br />

63<br />

<strong>CBS</strong> 115683<br />

<strong>CBS</strong> 366.80<br />

<strong>CBS</strong> 134.31<br />

<strong>CBS</strong> 813.71<br />

<strong>CBS</strong> 110024<br />

CPC 12921<br />

CPC 12046<br />

<strong>CBS</strong> 572.78<br />

CPC 12045<br />

CPC 11840<br />

<strong>CBS</strong> 157.82<br />

CPC 12212<br />

CPC 12211<br />

CPC 12042<br />

CPC 11386<br />

<strong>CBS</strong> 161.55<br />

<strong>Cladosporium</strong> sp. CPC 12485<br />

74<br />

83<br />

58<br />

85<br />

<strong>CBS</strong> 102044<br />

CPC 12040<br />

<strong>Cladosporium</strong> sp. CPC 12484<br />

<strong>Cladosporium</strong> sp. <strong>CBS</strong> 172.52<br />

<strong>Cladosporium</strong> sp. <strong>CBS</strong> 113741<br />

<strong>Cladosporium</strong> antarcticum <strong>CBS</strong> 690.92<br />

<strong>Cladosporium</strong> sp. <strong>CBS</strong> 113742<br />

<strong>Cladosporium</strong> subinflatum CPC 12041<br />

<strong>Cladosporium</strong> pseudiridis <strong>CBS</strong> 116463<br />

<strong>Cladosporium</strong> sinuosum CPC 11839<br />

93<br />

100<br />

51<br />

66<br />

81<br />

94<br />

65<br />

<strong>CBS</strong> 107.20 <strong>Cladosporium</strong> iridis<br />

<strong>CBS</strong> 138.40<br />

<strong>Cladosporium</strong> herbaroides CPC 12052<br />

100<br />

64<br />

95<br />

71<br />

61<br />

99<br />

CPC 12751<br />

CPC 12753<br />

CPC 11817<br />

CPC 12758<br />

CPC 12759<br />

CPC 12054<br />

<strong>CBS</strong> 175.82<br />

<strong>CBS</strong> 223.31<br />

51 CPC 12757<br />

56 <strong>CBS</strong> 299.67<br />

CPC 12756<br />

67 CPC 12755<br />

CPC 12752<br />

CPC 11603<br />

<strong>CBS</strong> 300.49<br />

CPC 11600<br />

CPC 11604<br />

CPC 11602<br />

CPC 11601<br />

<strong>CBS</strong> 111.82<br />

CPC 12177<br />

CPC 12183<br />

100<br />

CPC 12181<br />

CPC 12180<br />

CPC 12179<br />

CPC 12178<br />

<strong>Cladosporium</strong> subtilissimum<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> spinulosum<br />

<strong>Cladosporium</strong> variabile<br />

<strong>Cladosporium</strong> macrocarpum<br />

<strong>Cladosporium</strong> herbarum<br />

Fig. 3. One of 40 equally most parsimonious trees obtained from a heuristic search with 100 r<strong>and</strong>om taxon additions of the combined sequence alignment (ITS, ACT, CAL, EF,<br />

HIS). <strong>The</strong> scale bar shows ten changes, <strong>and</strong> bootstrap support values from 1 000 replicates are shown at the nodes. Thickened lines indicate the strict consensus branches<br />

<strong>and</strong> strain numbers in bold represent ex-type sequences. <strong>The</strong> tree was rooted to sequences of Cercospora beticola strain CPC 11557 (GenBank accession numbers AY840527,<br />

AY840458, AY840425, AY840494, AY840392, respectively).<br />

www.studiesinmycology.org<br />

111


Schubert et al.<br />

Davidiella sp.<br />

Global (Gapcost:0%)<br />

ACT<br />

80<br />

85<br />

D. macrosporum D. tassiana<br />

D. macrocarpa<br />

D. macrocarpa<br />

90<br />

D. macrospora<br />

Davidiella sp.<br />

D. variabile<br />

D. Variabile<br />

D. tassiana<br />

D. allicina<br />

D. alliacea<br />

95<br />

100<br />

• . <strong>CBS</strong> 673.69<br />

• . CPC 11606<br />

• . <strong>CBS</strong> 109082<br />

• . CPC 11609<br />

<br />

. <strong>CBS</strong> 110024<br />

<br />

. <strong>CBS</strong> 157.82<br />

<br />

. <strong>CBS</strong> 161.55<br />

<br />

. <strong>CBS</strong> 572.78<br />

<br />

. CPC 11386<br />

<br />

. CPC 12042<br />

<br />

. CPC 12046<br />

<br />

. CPC 12211<br />

<br />

. CPC 12212<br />

<br />

. CPC 11840<br />

<br />

. <strong>CBS</strong> 188.54<br />

<br />

. <strong>CBS</strong> 366.80<br />

<br />

. CPC 12921<br />

<br />

. <strong>CBS</strong> 159.54<br />

<br />

. <strong>CBS</strong> 177.71<br />

<br />

. <strong>CBS</strong> 399.80<br />

<br />

. <strong>CBS</strong> 521.68<br />

<br />

. CPC 12139<br />

<br />

. <strong>CBS</strong> 115683<br />

<br />

. CPC12041<br />

<br />

. CPC 11839<br />

<br />

. <strong>CBS</strong> 134.31<br />

<br />

<br />

. <strong>CBS</strong> 813.71<br />

<br />

. CPC 12045<br />

<br />

. <strong>CBS</strong> 113753<br />

<br />

. <strong>CBS</strong> 113754<br />

<br />

. CPC 12044<br />

<br />

. CPC 12485<br />

<br />

. <strong>CBS</strong> 690.92<br />

<br />

. <strong>CBS</strong> 113742<br />

<br />

. <strong>CBS</strong> 113741<br />

<br />

. <strong>CBS</strong> 172.52<br />

<br />

. <strong>CBS</strong> 116463<br />

<br />

. <strong>CBS</strong> 102044<br />

<br />

. CPC 12040<br />

<br />

. <strong>CBS</strong> 289.49<br />

<br />

. <strong>CBS</strong> 842.91<br />

<br />

. <strong>CBS</strong> 843.91<br />

<br />

. <strong>CBS</strong> 113744<br />

<br />

. CPC 12484<br />

<br />

. CPC 12751<br />

<br />

. CPC 12753<br />

•<br />

. <strong>CBS</strong> 107.20<br />

•<br />

. <strong>CBS</strong> 138.40<br />

<br />

. CPC 11600<br />

<br />

. CPC 11601<br />

<br />

. CPC 11602<br />

<br />

. CPC 11604<br />

<br />

. <strong>CBS</strong> 111.82<br />

<br />

. <strong>CBS</strong> 300.49<br />

<br />

. CPC 11603<br />

•<br />

. <strong>CBS</strong> 175.82<br />

•<br />

. <strong>CBS</strong> 223.31<br />

•<br />

. <strong>CBS</strong> 299.67<br />

•<br />

. CPC 11817<br />

•<br />

. CPC 12054<br />

<br />

. CPC 12752<br />

<br />

. CPC 12755<br />

<br />

. CPC 12756<br />

•<br />

. CPC 12758<br />

•<br />

. CPC 12759<br />

•<br />

. CPC 12757<br />

<br />

. CPC 12177<br />

<br />

. CPC 12178<br />

<br />

. CPC 12179<br />

<br />

. CPC 12180<br />

<br />

. CPC 12181<br />

<br />

. CPC 12183<br />

•<br />

. CPC 12052<br />

<br />

. CPC 11813<br />

<br />

. CPC 12051<br />

<br />

. CPC 12053<br />

<br />

. CPC 12043<br />

<br />

. CPC 12047<br />

<br />

. <strong>CBS</strong> 290.49<br />

<strong>Cladosporium</strong> cladosporioides complex<br />

<strong>Cladosporium</strong> cladosporioides complex<br />

<strong>Cladosporium</strong> cladosporioides complex<br />

<strong>Cladosporium</strong> cladosporioides complex<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> subinflatum<br />

<strong>Cladosporium</strong> sinuosum<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>Cladosporium</strong> subtilissimum s.str.<br />

<strong>Cladosporium</strong> subtilissimum s.str.<br />

<strong>Cladosporium</strong> subtilissimum s.str.<br />

<strong>Cladosporium</strong> subtilissimum s.lat.<br />

<strong>Cladosporium</strong> antarcticum<br />

<strong>Cladosporium</strong> subtilissimum s.lat.<br />

<strong>Cladosporium</strong> subtilissimum s.lat.<br />

<strong>Cladosporium</strong> subtilissimum s.lat.<br />

<strong>Cladosporium</strong> pseudiridis<br />

<strong>Cladosporium</strong> spinulosum<br />

<strong>Cladosporium</strong> spinulosum<br />

Davidiella sp.<br />

<strong>Cladosporium</strong> ossifragi<br />

<strong>Cladosporium</strong> ossifragi<br />

<strong>Cladosporium</strong> subtilissimum s.lat.<br />

<strong>Cladosporium</strong> subtilissimum s.lat.<br />

<strong>Cladosporium</strong> variabile<br />

<strong>Cladosporium</strong> variabile<br />

<strong>Cladosporium</strong> iridis<br />

<strong>Cladosporium</strong> iridis<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> macrocarpum<br />

<strong>Cladosporium</strong> macrocarpum<br />

<strong>Cladosporium</strong> macrocarpum<br />

<strong>Cladosporium</strong> macrocarpum<br />

<strong>Cladosporium</strong> macrocarpum<br />

<strong>Cladosporium</strong> macrocarpum<br />

<strong>Cladosporium</strong> macrocarpum<br />

<strong>Cladosporium</strong> macrocarpum<br />

<strong>Cladosporium</strong> macrocarpum<br />

<strong>Cladosporium</strong> macrocarpum<br />

<strong>Cladosporium</strong> macrocarpum<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> herbarum<br />

<strong>Cladosporium</strong> herbaroides<br />

<strong>Cladosporium</strong> tenellum<br />

<strong>Cladosporium</strong> tenellum<br />

<strong>Cladosporium</strong> tenellum<br />

<strong>Cladosporium</strong> ramotenellum<br />

<strong>Cladosporium</strong> ramotenellum<br />

Davidiella sp.<br />

Water<br />

Germany<br />

Wood<br />

Belgium<br />

Human<br />

Netherl<strong>and</strong>s<br />

Fungus<br />

Russia<br />

Tree<br />

Germany<br />

Hypersaline Slovenia<br />

Airco<br />

Slovenia<br />

Grass<br />

Belgium<br />

Grass<br />

Belgium<br />

Plant<br />

Australia<br />

? ?<br />

Human<br />

Netherl<strong>and</strong>s<br />

Tree<br />

Australia<br />

Human<br />

Netherl<strong>and</strong>s<br />

Chemical Netherl<strong>and</strong>s<br />

Human<br />

Netherl<strong>and</strong>s<br />

Air<br />

Netherl<strong>and</strong>s<br />

Grass<br />

Netherl<strong>and</strong>s<br />

Treated pole USA<br />

Hypersaline Slovenia<br />

Plant<br />

New Zeal<strong>and</strong><br />

? Germany<br />

Plant<br />

Czechia<br />

Hypersaline Spain<br />

Cherry<br />

USA .<br />

Grape<br />

USA<br />

Hypersaline Slovenia<br />

.<br />

Wood<br />

Argentina<br />

Lichen<br />

Antarctica<br />

Grape<br />

USA<br />

Grape<br />

USA<br />

Plant<br />

USA<br />

Plant<br />

New Zeal<strong>and</strong><br />

Hypersaline Slovenia<br />

Hypersaline Slovenia<br />

Plant<br />

Switzerl<strong>and</strong><br />

Plant<br />

Norway<br />

Plant<br />

Norway<br />

Grape<br />

USA<br />

Wood<br />

Argentina<br />

Spinach<br />

USA<br />

Spinach<br />

USA<br />

Plant<br />

Netherl<strong>and</strong>s<br />

Plant<br />

Netherl<strong>and</strong>s<br />

Plant<br />

USA<br />

Plant<br />

USA<br />

Plant<br />

USA<br />

Plant<br />

USA<br />

Plant<br />

Switzerl<strong>and</strong><br />

Plant<br />

Switzerl<strong>and</strong><br />

Plant<br />

USA<br />

Water<br />

Romania<br />

? ?<br />

Straw<br />

Turkey<br />

Branch<br />

USA<br />

Hypersaline Slovenia<br />

Spinach<br />

USA<br />

Spinach<br />

USA<br />

Spinach<br />

USA<br />

Spinach<br />

USA<br />

Spinach<br />

USA<br />

Spinach<br />

USA<br />

Grass<br />

Netherl<strong>and</strong>s<br />

Grass<br />

Netherl<strong>and</strong>s<br />

Grass<br />

Netherl<strong>and</strong>s<br />

Grass<br />

Netherl<strong>and</strong>s<br />

Grass<br />

Netherl<strong>and</strong>s<br />

Grass<br />

Netherl<strong>and</strong>s<br />

Saltern<br />

Israel<br />

Fungus<br />

USA<br />

Hypersaline Slovenia<br />

Hypersaline Slovenia<br />

Hypersaline Slovenia<br />

Airco<br />

Slovenia<br />

Plant<br />

Switzerl<strong>and</strong><br />

Fig. 4. Distance tree of the <strong>Cladosporium</strong> herbarum complex based on ACT sequence data generated with UPGMA, showing Structure analysis at K = 6 under admixture model<br />

with correlated allele frequencies. Group indications (18) are taken from a tree based on EF sequences with AIC under the HKYG model.<br />

112


<strong>Cladosporium</strong> herbarum species complex<br />

B<br />

A<br />

C<br />

D<br />

Fig. 5. Split decomposition of the <strong>Cladosporium</strong> herbarum complex using SplitsTree of 16–22 unique alleles obtained from 79 <strong>Cladosporium</strong> isolates for four loci. <strong>The</strong> star-like<br />

structures suggest clonal development. A = ACT, B = CAL, C = HIS, D = EF. Scale bars = 0.01 nucleotide substitutions per site.<br />

Taxonomy<br />

Key to the <strong>Cladosporium</strong> species treated<br />

Morphological features used in the key to distinguish the species treated in this study were determined after 7 d growth at 25 ºC on SNA using<br />

light microscopy, <strong>and</strong> cultural characteristics after 14 d incubation on PDA.<br />

1. Conidia usually smooth, rarely minutely verruculose ..................................................................... C. cladosporioides (species complex)<br />

1. Conidia with different surface ornamentation, minutely to distinctly verruculose, verrucose to echinulate or spiny<br />

.................................................................................................................................................................................................................... 2<br />

2. Conidiophores uniform, macronematous; conidia solitary, sometimes formed in short unbranched chains .............................................. 3<br />

2. Conidiophores both macronematous <strong>and</strong> micronematous; conidia always catenate, usually formed in branched chains<br />

.................................................................................................................................................................................................................... 5<br />

www.studiesinmycology.org<br />

113


Schubert et al.<br />

3. Conidiophores due to geniculations often growing zigzag-like, (4–)5–7 µm wide; conidia 9–21 × (5–)6–8 µm, 0–1-septate;<br />

conidiogenous loci <strong>and</strong> conidial hila 1.2–2(–2.2) µm diam .................................................................................................... C. sinuosum<br />

3. Conidiophores not growing zigzag-like, wider, 6–11 µm; conidia very large <strong>and</strong> wide, 15–75(–87) × (7–)10–19(–21) µm, often with more<br />

septa; conidiogenous loci <strong>and</strong> hila wider, (2–)2.5–4 µm diam ................................................................................................................... 4<br />

4. Conidia (18–)30–75(–87) × (7–)10–16(–18) µm, (0–)2–6(–7)-septate, walls thickened, especially in older conidia, up to 1 µm thick<br />

......................................................................................................................................................................................................... C. iridis<br />

4. Conidia shorter <strong>and</strong> wider, 15–55 × (9–)11–19(–21) µm, 0–3-septate, walls distinctly thickened, up to 2 µm, usually appearing zonate<br />

.............................................................................................................................................................................................. C. pseudiridis<br />

5(2) Macronematous conidiophores nodulose or nodose with conidiogenous loci usually confined to swellings ........................................... 6<br />

5. Macronematous conidiophores non-nodulose or only occasionally subnodulose due to geniculate proliferation, but conidiogenous loci not<br />

confined to swellings ................................................................................................................................................................................ 11<br />

6. Macronematous conidiophores 3–6 µm wide, swellings 5–11 µm wide .................................................................................................... 7<br />

6. Macronematous conidiophores somewhat narrower, (1.5–)2.5–5 µm wide, swellings 3–8 µm wide ........................................................ 8<br />

7. Aerial mycelium twisted; conidial septa often distinctly darkened, becoming sinuous with age, apex <strong>and</strong> base of the conidia often appear<br />

to be distinctly darkened; slower growing in culture (29 mm after 14 d on PDA) ...................................................................... C. variabile<br />

7. Aerial mycelium not twisted; conidial septa as well as apex <strong>and</strong> base not distinctly darkened, septa not sinuous with age; faster growing<br />

in culture (on average 38 mm after 14 d on PDA) .......................................................................................................... C. macrocarpum<br />

8. Macronematous conidiophores (1.5–)2.5–4.5(–5.5) µm wide, swellings 3–6.5 µm wide; conidia 4–17(–22) µm long, ornamentation<br />

variable, but usually densely echinulate, spines up to 0.8 µm long .................................................................................... C. subinflatum<br />

8. Macronematous conidiophores slightly wider, 3–5 µm, swellings (4–)5–8(–9) µm wide; conidia longer, up to 25(–35) µm, ornamentation<br />

minutely verruculose to verrucose, but not echinulate or spiny ................................................................................................................. 9<br />

9. Conidia formed by macronematous conidiophores 3–33 × (2–)3–6(–7) µm, with age becoming wider, (3.5–)5–9(–11) µm, darker <strong>and</strong><br />

more thick-walled ................................................................................................................................................................ C. herbaroides<br />

9. Conidia formed by macronematous conidiophores not becoming wider <strong>and</strong> darker with age, usually up to 7 µm wide<br />

.................................................................................................................................................................................................................. 10<br />

10. Conidiophores usually with small head-like swellings, sometimes also with a second intercalary nodule; small terminal conidia<br />

4–9 × 2.5–3.5 µm, secondary ramoconidia <strong>and</strong> occasionally formed ramoconidia 10–24(–31) × 3–5(–7) µm ........................ C. bruhnei<br />

10. Conidiophores with a single or often numerous swellings in short succession giving the stalk a knotty/gnarled appearance;<br />

conidia wider, small terminal conidia 4–10 × 3–5(–6) µm, intercalary conidia 6–16 × 4–6 µm, secondary ramoconidia 12–25(–35) ×<br />

(3–)5–7(–9) µm ....................................................................................................................................................................... C. herbarum<br />

11(5) Small terminal <strong>and</strong> intercalary conidia 4–15 × 3–5 µm, secondary ramoconidia 16–36(–40) × (4–)5–8 µm, 0–3(–4)-septate,<br />

ramoconidia absent ................................................................................................................................................................. C. ossifragi<br />

11. Small terminal conidia, ramoconidia <strong>and</strong> secondary ramoconidia distinctly narrower, 2–5(–6) µm wide, 0–2(–3)-septate<br />

.................................................................................................................................................................................................................. 12<br />

12. Mycelium dimorphic, narrow hyphae 1–3 µm wide, hyaline to subhyaline, thin-walled, hyphae of the second type wider, 3.5–8(–9)<br />

µm, pale to dark greyish olivaceous or olivaceous-brown, thick-walled, sometimes even two-layered, 1(–1.5) µm thick, hyphae<br />

appearing consistently enveloped in polysaccharide-like material or covered by a slime coat; conidiophores usually several times slightly<br />

to distinctly geniculate towards the apex, with numerous conidiogenous loci crowded towards the apex, up to 14 per conidiogenous cell<br />

............................................................................................................................................................................................. C. antarcticum<br />

12. Mycelium not dimorphic, neither enveloped in polysaccharide-like material nor covered by a slime coat; conidiophores usually not<br />

geniculate, occasionally only slightly so ................................................................................................................................................... 13<br />

13. Conidial ornamentation distinctly echinulate, spiny (baculate, digitate or capitate under SEM), spines 0.5–1.3 µm long, loose to moderately<br />

dense, conidial hila usually situated on small peg-like prolongations or denticles .............................................................. C. spinulosum<br />

13. Conidial ornamentation different, minutely verruculose to verruculose, conidial hila not situated on peg-like prolongations<br />

.................................................................................................................................................................................................................. 14<br />

14. Small terminal conidia narrowly obovoid, limoniform or fusiform, but neither globose nor subglobose; conidiogenous loci <strong>and</strong> conidial hila<br />

0.5–2(–2.5) µm diam ........................................................................................................................ C. subtilissimum (species complex)<br />

14. Numerous small globose or subglobose terminal conidia formed, also ovoid or limoniform; conidiogenous loci <strong>and</strong> conidial hila somewhat<br />

smaller, 0.5–1.5(–2) µm diam .................................................................................................................................................................. 15<br />

114


<strong>Cladosporium</strong> herbarum species complex<br />

15. Conidiophores usually with numerous conidiogenous loci forming sympodial clusters of pronounced scars at the apex, sometimes up to<br />

10 or even more denticulate loci; conidia 3–20(–28) × 2.5–5(–6) µm, 0–1(–2)-septate, often with several apically crowded hila, up to 7(–9)<br />

.................................................................................................................................................................................................. C. tenellum<br />

15. Conidiophores usually only with few conidiogenous loci, mostly 1–3; conidia longer <strong>and</strong> narrower, 2.5–35 × 2–4(–5) µm, 0–3-septate,<br />

usually with up to three distal conidial hila ....................................................................................................................... C. ramotenellum<br />

Key to the Davidiella species treated<br />

1. Ascospores frequently wider than 7 µm when mounted in Shear’s solution or lactic acid, apical cell obtusely rounded<br />

.................................................................................................................................................................................................................... 2<br />

1. Ascospores not wider than 7 µm when mounted in Shear’s solution or lactic acid, apical cell acutely rounded, ascospores (20–)25–27<br />

(–30) × (5.5–)6–7 µm ................................................................................................................................................................. D. allicina<br />

2. Pseudoparaphyses prominent; asci frequently >95 µm; ascospores (22–)23–26(–28) × (6–)6.5–7(–8) µm...................... D. macrocarpa<br />

2. Pseudoparaphyses mostly absent in older ascomata; asci


Schubert et al.<br />

Fig. 6. <strong>Cladosporium</strong> antarcticum (<strong>CBS</strong> 690.92). Macro- <strong>and</strong> micronematous conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

geniculate with several conidiogenous loci at the apex, 2–22 ×<br />

2–3 µm, pale greyish olivaceous, loci denticulate. Ramoconidia<br />

occasionally occurring, cylindrical, up to 30 µm long, 4–5 µm<br />

wide, 0–1-septate, concolorous with the tips of conidiophores, with<br />

a broadly truncate, unthickened <strong>and</strong> not darkened base, without<br />

dome <strong>and</strong> rim, 2.5 µm wide. Conidia catenate, in branched chains,<br />

straight, small terminal conidia obovoid, limoniform or narrowly<br />

ellipsoid, 4–14 × 2.5–4 µm [av. ± SD, 8.5 (± 3.3) × 3.5 (± 0.6)],<br />

0(–1)-septate, secondary ramoconidia ellipsoid to cylindrical, often<br />

with several or numerous conidial hila crowded at the distal end, up<br />

to 12, 13–30 × 4–5 µm [av. ± SD, 20.1 (± 5.8) × 4.3 (± 0.5) µm],<br />

0–3-septate, sometimes slightly constricted at the median septum,<br />

pale olivaceous-brown or greyish brown, minutely verruculose to<br />

verrucose (granulate under SEM), walls more or less thickened,<br />

rounded or slightly attenuated towards apex <strong>and</strong> base, hila<br />

protuberant, denticulate, 0.8–1.5(–2) µm diam, thickened <strong>and</strong><br />

darkened-refractive; microcyclic conidiogenesis occurring.<br />

Cultural characteristics: Colonies on PDA attaining 9 mm diam after<br />

14 d at 25 ºC, greenish olivaceous to grey-olivaceous, at the margin<br />

becoming dull green, reverse with a pale olivaceous-grey centre<br />

<strong>and</strong> a broad olivaceous-black margin, margin narrow, regular,<br />

entire edge, white, feathery, aerial mycelium sparse but colonies<br />

appearing felty, growth flat with somewhat elevated colony centre,<br />

prominent exudates not formed, sporulation dense, covering almost<br />

the whole colony. Colonies on MEA attaining 12 mm diam after 14<br />

d at 25 ºC, olivaceous-grey to iron-grey, iron-grey reverse, velvety<br />

to powdery, aerial mycelium sparse, sporulation profuse. Colonies<br />

on OA attaining 4 mm after 14 d at 25 ºC, olivaceous-grey, aerial<br />

mycelium sparse, diffuse, growth flat, without prominent exudates,<br />

sporulating.<br />

Specimen examined: Antarctica, King George, Arctowski, isolated from the lichen<br />

Caloplaca regalis (Teloschistaceae), C. Möller, No. 32/12, 1991, <strong>CBS</strong>-H 19857,<br />

holotype, isotype HAL 2024 F, culture ex-type <strong>CBS</strong> 690.92.<br />

Substrate <strong>and</strong> distribution: On the lichen Caloplaca regalis;<br />

Antarctica.<br />

Notes: This is the second genuine lichenicolous species of the<br />

<strong>genus</strong> <strong>Cladosporium</strong>. <strong>Cladosporium</strong> licheniphilum Heuchert & U.<br />

Braun, occurring on apothecia of Pertusaria alpina in Russia, is<br />

quite distinct from C. antarcticum by having subcylindrical or only<br />

slightly geniculate-sinuous, wider conidiophores, 5–8 µm, with<br />

numerous characteristic terminal branches <strong>and</strong> much shorter, 0–<br />

1-septate, smooth conidia, 3.5–13 × 3–7 µm (Heuchert & Braun<br />

2006). <strong>Cladosporium</strong> lichenicola Linds. was invalidly published <strong>and</strong><br />

C. arthoniae M.S. Christ. & D. Hawksw. as well as C. lichenum<br />

Keissl. are to be excluded from the <strong>genus</strong> <strong>Cladosporium</strong> since<br />

they do not possess the typical cladosporioid scar structure but<br />

inconspicuous, unthickened conidiogenous loci <strong>and</strong> conidial hila<br />

(Hawksworth 1979, Heuchert et al. 2005). <strong>The</strong> fungicolous species<br />

C. uredinicola Speg. <strong>and</strong> the foliicolous species C. alneum Pass.<br />

ex K. Schub. <strong>and</strong> C. psoraleae M.B. Ellis are morphologically<br />

superficially <strong>similar</strong>. However, C. uredinicola, a widespread fungus<br />

on rust fungi, downy mildews <strong>and</strong> powdery mildew fungi, differs<br />

in having somewhat longer <strong>and</strong> wider, smooth conidia, 3–39 × 2–<br />

6.5(–8) µm, <strong>and</strong> wider conidiogenous loci <strong>and</strong> conidial hila, 0.5–3<br />

116


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 7. <strong>Cladosporium</strong> antarcticum (<strong>CBS</strong> 690.92). A.<br />

Overview of the growth pattern on SNA. Note the very<br />

large bulbous cells formed at the base of different<br />

conidiophores. Other conidiophores sprout from the<br />

agar surface. B. Overview of conidiophores <strong>and</strong> conidia.<br />

Note the large distance of the scars on the conidiophore<br />

<strong>and</strong> the different stages of conidial formation on the<br />

tips of other conidia. <strong>The</strong> long secondary ramoconidia<br />

are also visible, <strong>and</strong> sparse aerial hyphae. C. Detail<br />

of B with details of the ornamentation <strong>and</strong> scars. <strong>The</strong><br />

absence of ornamentation at the apical (spore-forming)<br />

end of the secondary ramoconidium is clearly visible. D–<br />

E. Tubular structures on coniophore (D) <strong>and</strong> secondary<br />

ramoconidium (E). Scale bars: A–B = 10 μm, C–D = 5<br />

µm, E = 2 µm.<br />

Fig. 8. <strong>Cladosporium</strong> antarcticum (<strong>CBS</strong> 690.92). A–B. Macronematous conidiophores. C, G. Mycelium enveloped by a polysaccharide-like layer. D, F. Conidia. E. Micronematous<br />

conidiophore. H. Ramoconidium with numerous distal scars. Scale bars = 10 µm.<br />

µm (Heuchert et al. 2005); C. alneum, which causes leaf spots on<br />

Alnus glutinosa, possesses longer <strong>and</strong> wider conidiophores, 25–<br />

260 × (2–)3–7(–8.5) µm, <strong>and</strong> somewhat shorter, smooth conidia<br />

(Schubert 2005, Schubert et al. 2006); <strong>and</strong> C. psoraleae, known<br />

from Myanmar on Psoralea corylifolia, can easily be distinguished<br />

from C. antarcticum by its smooth <strong>and</strong> wider conidia, 3.5–7 µm,<br />

<strong>and</strong> wider conidiogenous loci <strong>and</strong> conidial hila, 1–3 µm diam (Ellis<br />

1972, Schubert 2005).<br />

www.studiesinmycology.org<br />

117


Schubert et al.<br />

Fig. 9. <strong>Cladosporium</strong> bruhnei (CPC 12211). Macro- <strong>and</strong> micronematous conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

<strong>Cladosporium</strong> bruhnei Linder, Bull. Natl. Mus. Canada 97: 259.<br />

1947. Figs 9–12.<br />

≡ Hormodendrum hordei Bruhne, in W. Zopf, Beitr. Physiol. Morph. nied.<br />

Org. 4: 1. 1894, non C. hordei Pass., 1887.<br />

≡ <strong>Cladosporium</strong> herbarum (Pers.: Fr.) Link var. (δ) cerealium Sacc. f.<br />

hordei (Bruhne) Ferraris, Flora Ital. Crypt., Pars I, Fungi, Fasc.13: 882.<br />

1914.<br />

≡ <strong>Cladosporium</strong> hordei (Bruhne) Pidopl., Gribnaja Flora Grubych Kormov:<br />

268. 1953, nom. illeg., homonym, non C. hordei Pass., 1887.<br />

Teleomorph: Davidiella allicina (Fr. : Fr.) Crous & Aptroot, in<br />

Aptroot, Mycosphaerella <strong>and</strong> its anamorphs: 2. Conspectus of<br />

Mycosphaerella. <strong>CBS</strong> Biodiversity Ser. 5: 30. 2006.<br />

Basionym: Sphaeria allicina Fr., Kongl. Vetensk. Acad. H<strong>and</strong>l. 38:<br />

247. 1817, sactioned by Fr., Syst. Mycol. 2: 437. 1823.<br />

≡ Sphaerella allicina (Fr. : Fr.) Auersw., in Gonn. & Rabenh., Mycol.<br />

Europaea 5–6: 19. 1869.<br />

Ascomata pseudothecial, black, superficial, situated on a small<br />

stroma, globose, up to 250 µm diam; ostioles periphysate, with<br />

apical periphysoids present; wall consisting of 3–6 layers of<br />

reddish brown textura angularis. Asci fasciculate, bitunicate,<br />

subsessile, obovoid to broadly ellipsoid, straight to slightly curved,<br />

8-spored, 65–90 × 16–25 µm; with pseudoparenchymatal cells<br />

of the hamathecium persistent. Ascospores tri- to multiseriate,<br />

overlapping, hyaline, with irregular lumina, thick-walled, straight<br />

to slightly curved, fusoid-ellipsoidal with obtuse basal end, <strong>and</strong><br />

acutely rounded apical end, widest near the middle of the apical<br />

cell, medianly 1-septate, not to slightly constricted at the septum,<br />

(20–)25–27(–30) × (5.5–)6–7 µm.<br />

Mycelium superficial, hyphae branched, 1.5–8 µm wide,<br />

pluriseptate, broader hyphae usually slightly constricted at the<br />

septa <strong>and</strong> somewhat swollen, hyaline to subhyaline, almost smooth<br />

to somewhat verruculose or irregularly rough-walled, sometimes<br />

appearing to have a slime coat, walls unthickened. Conidiophores<br />

macronematous, sometimes also micronematous, arising as lateral<br />

or terminal branches from plagiotropous or ascending hyphae, erect,<br />

straight to more or less flexuous, sometimes geniculate, nodulose,<br />

usually with small head-like swellings, sometimes also with<br />

intercalary nodules, sometimes swellings protruding <strong>and</strong> elongated<br />

to one side, unbranched, occasionally branched, (7–)20–330 µm,<br />

sometimes even longer, (2–)3–5 µm wide, swellings (4–)5–8 µm<br />

wide, pluriseptate, not constricted at the septa, septa sometimes<br />

not very conspicuous, subhyaline to pale brown or pale olivaceous,<br />

smooth or somewhat verruculose, walls unthickened or almost so,<br />

more thickened with age. Conidiogenous cells integrated, usually<br />

terminal, cylindrical with a terminal head-like swelling, sometimes<br />

with a second swelling, 15–40 µm long, proliferation sympodial,<br />

with few conidiogenous loci confined to swellings, up to five per<br />

swelling, loci protuberant, conspicuous, 1–2 µm diam, thickened<br />

<strong>and</strong> darkened-refractive. Conidia catenate, formed in branched<br />

chains, straight to slightly curved, small terminal conidia subglobose,<br />

ovoid to obovoid or somewhat limoniform, 4–9 × 2.5–3.5 µm [av. ±<br />

SD, 6.5 (± 1.5) × 3.1 (± 0.5) µm], aseptate; secondary ramoconidia<br />

<strong>and</strong> occasionally formed ramoconidia ellipsoid to subcylindrical<br />

or cylindrical, 10–24(–31) × 3–5(–7) µm [av. ± SD, 16.1 (± 4.1)<br />

× 4.1 (± 0.8) µm], rarely up to 40 µm long, 0–1(–3)-septate, very<br />

rarely 5-septate, subhyaline to pale brown or pale olivaceous,<br />

minutely verruculose to verrucose (mostly granulate with some<br />

muricate projections under SEM), walls unthickened or almost so,<br />

apex rounded or slightly attenuated towards apex <strong>and</strong> base, hila<br />

protuberant, conspicuous, 1–2 µm wide, up to 1 µm high, thickened<br />

<strong>and</strong> darkened-refractive; microcyclic conidiogenesis occurring.<br />

118


<strong>Cladosporium</strong> herbarum species complex<br />

Belgium, isolated from Quercus robur (Fagaceae), <strong>CBS</strong> 157.82; Kampenhout,<br />

isolated from Hordeum vulgare (Poaceae), 26 June 2005, J.Z. Groenewald, <strong>CBS</strong>-H<br />

19856, neotype designated here of C. bruhnei, isoneotype HAL 2023 F, cultures<br />

ex-type <strong>CBS</strong> 121624 = CPC 12211, CPC 12212. Czech Republic, Lisen, isolated<br />

from Polygonatum odoratum (Liliaceae), <strong>CBS</strong> 813.71, albino mutant of <strong>CBS</strong> 812.71.<br />

Germany, <strong>CBS</strong> 134.31 = ATCC 11283 = IMI 049632; Nordrhein-Westfalen, Mühlheim<br />

an der Ruhr, isolated from industrial water, IWW 727, <strong>CBS</strong> 110024; Sachsen-Anhalt,<br />

Halle (Saale), Robert-Franz-Ring, isolated from leaves of Tilia cordata (Tiliaceae),<br />

2004, K. Schubert, CPC 11386. Netherl<strong>and</strong>s, isolated from air, <strong>CBS</strong> 521.68; isolated<br />

from Hordeum vulgare, 1 Jan. 2005, P.W. Crous, CPC 12139; isolated from man,<br />

skin, <strong>CBS</strong> 159.54 = ATCC 36948; Amsterdam, isolated from Thuja tincture, <strong>CBS</strong><br />

177.71; Geleen, St. Barbara Ziekenhuis, isolated from man, skin, <strong>CBS</strong> 366.80, <strong>CBS</strong><br />

399.80; isolated from man, sputum, Aug. 1955, <strong>CBS</strong> 161.55. New Zeal<strong>and</strong>, Otago,<br />

Lake Harris, isolated from Ourisia macrophylla (Scrophulariaceae), 30 Jan. 2005,<br />

A. Blouin, Hill 1135, CPC 11840. Russia, Moscow region, isolated from Polyporus<br />

radiatus (Polyporaceae), Oct. 1978, <strong>CBS</strong> 572.78 = VKM F-405. Slovenia, Ljubljana,<br />

isolated from an air conditioning system, 2004, M. Butala, EXF-680 = CPC 12046;<br />

Sečovlje, isolated from hypersaline water from salterns (reserve pond), 2005, P.<br />

Zalar, EXF-389 = CPC 12042. Spain, Ebro Delta, isolated from hypersaline water<br />

from salterns (crystallisation pond), 2004, P. Zalar, EXF-594 = CPC 12045. Sweden,<br />

Skåne, on tip blight of living leaves of Allium sp. (Alliaceae), Fr. no. F-09810, UPS-<br />

FRIES, holotype of Davidiella allicina. U.S.A., New York, Geneva, isolated from<br />

CCA-treated Douglas-fir pole, <strong>CBS</strong> 115683 = ATCC 66670 = CPC 5101.<br />

Substrate <strong>and</strong> distribution: Living <strong>and</strong> decaying plant material, man,<br />

air, hypersaline <strong>and</strong> industrial water; widespread.<br />

Literature: Saccardo (1899: 1076), Linder (1947: 289).<br />

Fig. 10. <strong>Cladosporium</strong> bruhnei (CPC 12211). A. Conidiophore with characteristic<br />

long secondary ramoconidium <strong>and</strong> complex conidiophore. B. Detail of hila on<br />

secondary ramoconidia. C. Details of prominent ornamentation on conidia. Scale<br />

bars: A = 10 µm, B = 2 µm, C = 5 µm.<br />

Cultural characteristics: Colonies on PDA reaching 22–32 mm diam<br />

after 14 d at 25 ºC, olivaceous-grey to iron-grey, sometimes whitish,<br />

smoke-grey to pale olivaceous due to abundant aerial mycelium<br />

covering almost the whole colony, with age collapsing becoming<br />

olivaceous-grey, occasionally zonate, velvety to floccose, margin<br />

narrow, entire edge, white, glabrous to somewhat feathery, aerial<br />

mycelium sparse to abundant, white, fluffy, growth regular, flat to<br />

low convex, sometimes forming few exudates in the colony centre,<br />

sporulating. Colonies on MEA reaching 21–32 mm diam after 14<br />

d at 25 ºC, grey-olivaceous, olivaceous-grey to dull green or irongrey,<br />

sometimes whitish to pale smoke-grey due to abundant aerial<br />

mycelium, olivaceous-grey to iron-grey reverse, velvety, margin<br />

narrow, entire edge to slightly undulate, white, radially furrowed,<br />

glabrous to slightly feathery, aerial mycelium sparse to abundant,<br />

mainly in the centre, white, fluffy, growth convex to raised, radially<br />

furrowed, distinctly wrinkled in the colony centre, without prominent<br />

exudates, sporulating. Colonies on OA reaching 20–32 mm diam<br />

after 14 d at 25 ºC, smoke-grey, grey-olivaceous to olivaceous-grey,<br />

greenish black or iron-grey reverse, margin narrow, entire edge,<br />

colourless to white, glabrous, aerial mycelium sparse to abundant,<br />

dark smoke-grey, diffuse, high, later collapsed, felty, growth flat,<br />

without prominent exudates, sporulation profuse.<br />

Specimens examined: Sine loco et dato, <strong>CBS</strong> 188.54 = ATCC 11290 = IMI<br />

049638. Australia, N.S.W., Barrington Tops National Park, isolated from leaves<br />

of Eucalyptus stellulata (Myrtaceae), 3 Jan. 2006, B. Summerell, CPC 12921.<br />

Fig. 11. Davidiella allicina (F-09810, UPS-FRIES, holotype). Ascus <strong>and</strong> ascospores.<br />

Scale bar = 10 µm. P.W. Crous del.<br />

www.studiesinmycology.org<br />

119


Schubert et al.<br />

Fig. 12. <strong>Cladosporium</strong> bruhnei (CPC 12211) <strong>and</strong> its teleomorph Davidiella allicina. A–B. Macronematous conidiophores. C. Conidial chains. D. Micronematous conidiophore. E.<br />

Ascomata of the teleomorph formed on the host. F–G. Asci. Scale bars: A–B, D, F = 10 µm, E = 200 µm.<br />

Notes: <strong>Cladosporium</strong> bruhnei proved to be an additional component<br />

of the herbarum complex. <strong>The</strong> species resembles C. herbarum s.<br />

str. as already stated by Linder (1947), but possesses consistently<br />

narrower conidia, usually 2.5–5 µm wide, <strong>and</strong> the conidiophores<br />

often form only a single apical swelling. <strong>The</strong> species was described<br />

by Bruhne (l.c.) as Hormodendrum hordei from Germany but type<br />

material could not be located. Linder (1947) examined No. 1481a-5<br />

(Canada, N. Quebec, Sugluk, on Elymus arenarius var. villosus,<br />

31 Jul. 1936, E. Meyer), presumably in the National Museum, <strong>and</strong><br />

stated that this specimen agreed well with the description <strong>and</strong><br />

illustration given by Bruhne (l.c.). Although the species occurs<br />

on numerous substrates <strong>and</strong> is widely distributed, it has not yet<br />

been recognised as a distinct species since it has probably been<br />

interpreted as a narrow variant of C. herbarum.<br />

Based on morphology <strong>and</strong> DNA sequence data, the <strong>CBS</strong> strain<br />

<strong>CBS</strong> 177.71 chosen by Prasil & de Hoog (1988) as representative<br />

living strain of C. herbarum, rather clusters together with isolates<br />

of C. bruhnei. <strong>The</strong> strain <strong>CBS</strong> 813.71 is an albino mutant of the<br />

latter species as it does not appear to contain colour pigment.<br />

Furthermore, all isolates from humans treated until now as C.<br />

herbarum proved to be conspecific with the narrow-spored C.<br />

bruhnei.<br />

Although Davidiella tassiana (ascospores 17–25 × 6–8.5 µm,<br />

RO) was treated as synonymous to D. allicina (ascospores 20–27<br />

× 6–7 µm, UPS) in Aptroot (2006), they differ in apical ascospore<br />

taper, with ascospores of D. allicina being acutely rounded, while<br />

those of D. tassiana are obtusely rounded. <strong>The</strong> same ascospore<br />

taper was also observed in the teleomorph of C. bruhnei, <strong>and</strong> thus<br />

the name D. allicina is herewith linked to C. bruhnei, which is distinct<br />

from C. herbarum, having D. tassiana as teleomorph.<br />

<strong>Cladosporium</strong> herbaroides K. Schub., Zalar, Crous & U. Braun,<br />

sp. nov. MycoBank MB504574. Figs 13–15.<br />

Etymology: Refers to its morphological <strong>similar</strong>ity to <strong>Cladosporium</strong><br />

herbarum.<br />

Differt a Cladosporio herbaro conidiis polymorphis, 3–33 × (2–)3–6(–7) µm,<br />

postremo latioribus, (3.5–)5–9(–11) µm, fuscis et crassitunicatis; et a Cladosporio<br />

macrocarpo conidiophoris leniter angustioribus, 3–5 µm latis, nodulis angustioribus,<br />

5–8 µm latis.<br />

Mycelium branched, (1–)2–8 µm wide, septate, often with small<br />

swellings <strong>and</strong> constrictions, subhyaline to pale brown or pale<br />

olivaceous-brown, smooth or almost so to somewhat verruculose,<br />

walls unthickened or almost so. Conidiophores macronematous<br />

<strong>and</strong> micronematous, arising lateral from plagiotropous hyphae or<br />

terminally from ascending hyphae. Macronematous conidiophores<br />

erect, straight to slightly flexuous, often geniculate, nodulose, with<br />

unilateral or multilateral swellings, often numerous swellings in<br />

short succession giving them a gnarled appearance, often forming<br />

somewhat protruding or prolonged lateral swellings or a branch-<br />

120


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 13. <strong>Cladosporium</strong> herbaroides (CPC 12052). Macro- <strong>and</strong> micronematous conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

like prolongation below the terminal swelling (due to sympodial<br />

proliferation), unbranched or sometimes branched, 30–230 µm<br />

long or even longer, 3–5 µm wide, swellings 5–8 µm wide, septate,<br />

not constricted at septa, pale to medium olivaceous-brown,<br />

smooth or almost so, walls slightly thickened. Conidiogenous cells<br />

integrated, terminal or intercalary, cylindrical, usually nodulose to<br />

nodose forming distinct swellings, sometimes geniculate, 15–55<br />

µm long, with numerous conidiogenous loci usually confined to<br />

swellings or situated on small lateral shoulders, sometimes on the<br />

top of short peg-like prolongations or denticles, loci protuberant,<br />

1–2 µm diam, thickened <strong>and</strong> darkened-refractive. Micronematous<br />

conidiophores much shorter, narrower, paler, neither nodulose nor<br />

geniculate, arising laterally from plagiotropous hyphae, often only<br />

as short lateral denticles or branchlets of hyphae, erect, straight,<br />

conical to cylindrical, unbranched, 3–65 × 2–3 µm, mostly aseptate,<br />

sometimes up to five septa, subhyaline, smooth, walls unthickened.<br />

Conidiogenous cells integrated, terminal or conidiophores reduced<br />

to conidiogenous cells, conidiogenous loci solitary or sometimes<br />

as sympodial clusters of pronounced denticles, protuberant, 1–1.5<br />

µm diam, thickened <strong>and</strong> somewhat darkened-refractive. Conidia<br />

polymorphous, two main morphological types recognisable, formed<br />

www.studiesinmycology.org<br />

by the two different types of conidiophores, conidia formed by<br />

macronematous conidiophores catenate, in branched chains,<br />

straight to slightly curved, subglobose, obovoid, limoniform, ellipsoid<br />

to cylindrical, 3–33 × (2–)3–6(–7) µm [av. ± SD, 14.5 (± 7.9) ×<br />

5.2 (± 1.2) µm], 0–2(–3)-septate, sometimes slightly constricted at<br />

septa, septa median or somewhat in the lower half, pale to medium<br />

olivaceous-brown, verruculose to verrucose (granulate under<br />

SEM), walls slightly thickened, with up to three rarely four distal<br />

scars, with age becoming medium or even dark brown (chocolate<br />

brown), wider <strong>and</strong> more thick-walled, 5.5–33 × (3.5–)5–9(–11) µm<br />

[av. ± SD, 14.4 (± 6.9) × 7.2 (± 1.9) µm], walls up to 1 µm thick,<br />

hila protuberant, 0.8–2(–2.5) µm diam, thickened <strong>and</strong> darkenedrefractive;<br />

microcyclic conidiogenesis occurring. Conidia formed by<br />

micronematous conidiophores paler <strong>and</strong> narrower, mostly formed<br />

in unbranched chains, sometimes in branched chains with up to<br />

three distal hila, straight to slightly curved, limoniform, narrowly<br />

fusiform, almost filiform to subcylindrical, 10–26(–35) × 2–3.5<br />

µm [av. ± SD, 15.6 (± 6.2) × 2.9 (± 0.5) µm], 0–1(–3)-septate,<br />

subhyaline to pale brown, almost smooth to minutely verruculose,<br />

walls unthickened, hila protuberant, 1–1.5 µm diam, thickened <strong>and</strong><br />

somewhat darkened-refractive.<br />

121


Schubert et al.<br />

Fig. 14. <strong>Cladosporium</strong> herbaroides (CPC 12052). A–B, D. Macronematous conidiophores. C. Conidial chain. E. Micronematous conidiophore. F. Microcyclic conidiogenesis. G.<br />

Conidia formed by micronematous conidiophores. Scale bars = 10 µm.<br />

Cultural characteristics: Colonies on PDA attaining 23 mm diam<br />

after 14 d at 25 ºC, grey-olivaceous to olivaceous, olivaceous-grey<br />

reverse, velvety, margin regular, entire edge, narrow, feathery, aerial<br />

mycelium abundantly formed, loose, with age covering large parts<br />

of the colony, woolly, growth flat with somewhat elevated colony<br />

centre, folded, regular, deep into the agar, with few prominent<br />

exudates, sporulation profuse. Colonies on MEA attaining 24 mm<br />

diam after 14 d at 25 ºC, grey- to greenish olivaceous, olivaceousgrey<br />

or iron-grey reverse, velvety to powdery, margin narrow,<br />

colourless, entire edge, somewhat feathery, aerial mycelium pale<br />

olivaceous-grey, sparse, growth convex, radially furrowed, folded<br />

in the colony centre, without prominent exudates, sporulating.<br />

Colonies on OA attaining 23 mm diam after 14 d at 25 ºC, greyolivaceous,<br />

margin more or less regular, entire edge, colourless,<br />

somewhat feathery, aerial mycelium whitish to smoke grey, at first<br />

sparse, later more abundantly formed, growth flat, without exudates,<br />

sporulation profuse.<br />

Specimen examined: Israel, from hypersaline water of Eilat salterns, 2004, coll.<br />

N. Gunde-Cimerman, isol. M. Ota, <strong>CBS</strong>-H 19858, holotype, isotype HAL 2025 F,<br />

culture ex-type <strong>CBS</strong> 121626 = EXF-1733 = CPC 12052.<br />

Substrate <strong>and</strong> distribution: Hypersaline water; Israel.<br />

Notes: <strong>Cladosporium</strong> herbaroides is morphologically <strong>similar</strong> to C.<br />

herbarum but differs in having somewhat longer conidia becoming<br />

wider, darker <strong>and</strong> even more thick-walled with age [at first conidia<br />

3–33 × (2–)3–6(–7) µm, with age (3.5–)5–9(–11) µm wide]. Besides<br />

that, the species often produces a second conidial type formed on<br />

micronematous conidiophores, giving rise to unbranched conidial<br />

chains which are almost filiform, limoniform, narrowly fusiform to<br />

subcylindrical, much narrower <strong>and</strong> paler than the ones formed by<br />

macronematous conidiophores, 10–26(–35) × 2–3.5 µm. In C.<br />

herbarum, conidia formed by micronematous conidiophores do not<br />

occur as frequently as in C. herbaroides, <strong>and</strong> differ in being often<br />

clavate <strong>and</strong> somewhat wider, up to 4(–5) µm wide. <strong>Cladosporium</strong><br />

macrocarpum is easily distinguishable by having somewhat wider<br />

conidiophores (3–)4–6 µm, with distinctly wider swellings, 5–10 µm<br />

wide, <strong>and</strong> the conidia are usually (3–)5–9(–10) µm wide.<br />

<strong>Cladosporium</strong> herbarum (Pers. : Fr.) Link, Ges. Naturf. Freunde<br />

Berlin Mag. Neuesten Entdeck. Gesammten Naturk. 7: 37. 1816:<br />

Fr., Syst. mycol. 3(2): 370. 1832. Figs 16–19.<br />

Basionym: Dematium herbarum Pers., Ann. Bot. (Usteri) 11: 32.<br />

1794: Fr., Syst. mycol. 3(2): 370. 1832.<br />

= Dematium epiphyllum var. (β) chionanthi Pers., Mycol. eur. 1: 16. 1822, syn.<br />

nov.<br />

For additional synonyms see Dugan et al. (2004), Schubert<br />

(2005).<br />

Teleomorph: Davidiella tassiana (De Not.) Crous & U. Braun,<br />

Mycol. Progr. 2: 8. 2003.<br />

Basionym: Sphaerella tassiana De Not., Sferiacei Italici 1: 87.<br />

1863.<br />

≡ Mycosphaerella tassiana (De Not.) Johanson, Öfvers. Förh. Kongl.<br />

Svenska Vetensk.-Akad. 41: 167. 1884.<br />

122


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 15. <strong>Cladosporium</strong> herbaroides (CPC 12052). A. Overview of the growth characteristics of this fungus. Broad hyphae run over the surface of the agar, <strong>and</strong> possibly give rise<br />

to conidiophore branches. <strong>The</strong> conidiophores of this fungus can be rather long, resembling aerial hyphae. Clusters of conidia are clearly visible in this micrograph. B. <strong>The</strong> very<br />

wide surface hyphae can anastomose. C. Conidiophore with secondary ramoconidia <strong>and</strong> conidia. Note the variation in scar size. D. A very elaborate, complex conidiophore with<br />

different scars of variable size, one being more than 2 μm wide! E. Details of secondary ramoconidia <strong>and</strong> hila. Note the rather strong ornamentation in which smaller “particles”<br />

are between larger ones. F. Three conidia in a row. Note the scar formation in the chain <strong>and</strong> the reduction of the size of the cells throughout the spore-chain. <strong>The</strong> inset shows<br />

the resemblance of the scars on a conidiophore <strong>and</strong> on a secondary ramoconidium. Scale bars: A = 50 µm, B–C, F (inset) = 10 µm, D–E = 5 µm, F = 2 µm.<br />

Ascomata pseudothecial, black, globose, erumpent to superficial,<br />

up to 200 µm diam, with 1(–3) short, periphysate ostiolar necks;<br />

wall consisting of 3–6 layers of medium red-brown textura<br />

angularis. Asci fasciculate, bitunicate, subsessile, obovoid to<br />

broadly ellipsoid, straight to slightly curved, 8-spored, 65–85 × 13–<br />

17 µm. Pseudoparaphyses absent in host material, but remnants<br />

observed when studied in culture, hyaline, septate, subcylindrical,<br />

anastomosing, 3–4 µm wide. Ascospores tri- to multiseriate,<br />

overlapping, hyaline, with irregular luminar inclusions, thickwalled,<br />

straight to slightly curved, fusoid-ellipsoidal with obtuse<br />

ends, widest near middle of apical cell, medianly 1-septate, not to<br />

slightly constricted at the septum, tapering towards both ends, but<br />

more prominently towards the lower end, (17–)20–23(–25) × (6–)<br />

7(–8) µm; becoming brown <strong>and</strong> verruculose in asci. Ascospores<br />

germinating after 24 h on MEA from both ends, with spore body<br />

becoming prominently constricted at the septum, but not distorting,<br />

up to 7 µm wide, hyaline to pale brown <strong>and</strong> appearing somewhat<br />

verruculose, enclosed in a mucoid sheath, with germ tubes being<br />

irregular, somewhat nodular.<br />

www.studiesinmycology.org<br />

123


Schubert et al.<br />

almost smooth to minutely verruculose or irregularly rough-walled,<br />

sometimes forming clavate conidia, up to 33 µm long, 0–2-septate.<br />

Conidiogenous cells integrated, terminal or conidiophores reduced<br />

to conidiogenous cells, narrowly cylindrical or filiform, with a single<br />

or two loci. Conidia catenate, in unbranched or loosely branched<br />

chains with branching mostly occurring in the lower part of the chain,<br />

straight to slightly curved, small terminal conidia without distal hilum<br />

obovoid, 4–10 × 3–5(–6) µm [av. ± SD, 7.8 (± 1.9) × 4.7 (± 0.9) µm],<br />

aseptate, intercalary conidia with a single or sometimes up to three<br />

distal hila limoniform, ellipsoid to subcylindrical, 6–16 × 4–6 µm<br />

[av. ± SD, 12.4 (± 1.6) × 5.3 (± 0.6) µm], 0–1-septate, secondary<br />

ramoconidia with up to four distal hila, ellipsoid to cylindrical-oblong,<br />

12–25(–35) × (3–)5–7(–9) µm [av. ± SD, 18.8 (± 4.5) × 6.2 (± 0.9)<br />

µm], 0–1(–2)-septate, rarely with up to three septa, sometimes<br />

distinctly constricted at the septum, septum median or somewhat<br />

in the upper or lower half, pale greyish brown or brown to medium<br />

brown or greyish brown, minutely verruculose to verrucose, walls<br />

slightly to distinctly thickened, guttulate to somewhat granular,<br />

usually only slightly attenuated towards apex <strong>and</strong> base, apex<br />

obtuse or slightly truncate, towards the base sometimes distinctly<br />

attenuated with hila situated on short stalk-like prolongations, hila<br />

slightly to distinctly protuberant, truncate to slightly convex, (0.8–)<br />

1–2.5(–3) µm wide, 0.5–1 µm high, somewhat thickened <strong>and</strong><br />

darkened-refractive; microcyclic conidiogenesis occurring, conidia<br />

forming micro- <strong>and</strong> macronematous secondary conidiophores.<br />

Fig. 16. Davidiella tassiana (RO, holotype). Ascus <strong>and</strong> ascospores. Scale bar = 10<br />

µm. P.W. Crous del.<br />

Mycelium superficial, loosely branched, (0.5–)1–5 µm wide, septate,<br />

sometimes constricted at septa, hyaline, subhyaline to pale brown,<br />

smooth or almost so to verruculose or irregularly rough-walled,<br />

sometimes appearing irregular in outline due to small swellings<br />

<strong>and</strong> constrictions, walls unthickened to somewhat thickened, cell<br />

lumen appearing to be granular. Conidiophores both macro- <strong>and</strong><br />

micronematous, arising laterally from plagiotropous hyphae or<br />

terminally from ascending hyphae. Macronematous conidiophores<br />

erect, straight to flexuous, somewhat geniculate-sinuous, nodulose<br />

to nodose with unilateral or multilateral swellings, with a single to<br />

numerous swellings in short succession giving the stalk a knotty/<br />

gnarled appearance, unbranched or occasionally branched, up to<br />

three times, sometimes with a lateral branch-like proliferation below<br />

or at the apex, 10–320 × 3.5–5 µm, swellings 5–8(–9) µm wide,<br />

pluriseptate, septa sometimes constricted when formed after a<br />

node, pale to medium brown, older ones almost dark brown, paler<br />

towards the apex, smooth or minutely verruculose, walls thickened,<br />

sometimes even two-layered. Conidiogenous cells integrated,<br />

terminal or intercalary, nodulose to nodose, with a single or up<br />

to five swellings per cell, 10–24 µm long, proliferation sympodial,<br />

with several conidiogenous loci confined to swellings, mostly<br />

situated on small lateral shoulders, more or less protuberant,<br />

broadly truncate to slightly convex, 1.5–2.5 µm diam, thickened<br />

<strong>and</strong> somewhat darkened-refractive. Micronematous conidiophores<br />

hardly distinguishable from hyphae, sometimes only as short lateral<br />

outgrowth with a single apical scar, short, conical to almost filiform<br />

or narrowly cylindrical, non-nodulose, not geniculate, unbranched,<br />

5–120 × 1.5–3(–4) µm, pluriseptate, not constricted at septa,<br />

cells usually very short, 5–15 µm long, subhyaline to pale brown,<br />

Cultural characteristics: Colonies on PDA reaching 19–37 mm diam<br />

after 14 d at 25 ºC, grey-olivaceous to olivaceous-grey, whitish<br />

to smoke-grey or pale olivaceous-grey due to abundant aerial<br />

mycelium, velvety, reverse olivaceous-grey or iron-grey, margin<br />

almost colourless, regular, entire edge, glabrous to feathery,<br />

aerial mycelium abundant mainly in the colony centre, dense,<br />

felty, woolly, sometimes becoming somewhat reddish brown,<br />

fawn coloured, growth regular, flat to low convex with an elevated<br />

colony centre, sometimes forming few large prominent exudates,<br />

sporulation profuse. Colonies on MEA reaching 17–37 mm diam<br />

after 14 d at 25 ºC, smoke-grey to pale olivaceous-grey towards<br />

margin, olivaceous-grey to iron-grey reverse, velvety, margin white,<br />

entire edge to slightly undulate, aerial mycelium abundant, dense,<br />

fluffy to felty, growth low convex or raised, radially furrowed, folded<br />

<strong>and</strong> wrinkled in the colony centre, without prominent exudates but<br />

sporulating. Colonies on OA reaching 12–28 mm diam after 14<br />

d at 25 ºC, olivaceous-grey to iron-grey, due to abundant aerial<br />

mycelium pale olivaceous-grey, olivaceous-grey reverse, margin<br />

narrow, more or less undulate, white, aerial mycelium white, loose to<br />

dense, high, fluffy to felty, covering large parts of the colony, growth<br />

flat to low convex, without prominent exudates, sporulating.<br />

Specimens examined: Sine loco, sine dato, L 910.225-733, lectotype of C.<br />

herbarum, selected by Prasil & de Hoog, 1988. Sine loco, on leaves of Chionanthus<br />

sp. (Oleaceae), L 910.255-872 = L-0115833, holotype of Dematium epiphyllum<br />

var. (β) chionanthi. Netherl<strong>and</strong>s, Wageningen, isolated from Hordeum vulgare<br />

(Poaceae), 2005, P.W. Crous, <strong>CBS</strong>-H 19853, epitype designated here of C.<br />

herbarum <strong>and</strong> D. tassiana, isoepitype HAL 2022 F, ex-type cultures, CPC 12177 =<br />

<strong>CBS</strong> 121621, CPC 12178–12179, 12181, 12183. Italy, on upper <strong>and</strong> lower surface<br />

of dead leaves of Carex nigra [“fusca”] (Cyperaceae), Tassi no. 862, RO, holotype<br />

of Davidiella tassiana. U.S.A., Colorado, San Juan Co., above Little Molas Lake,<br />

isolated from stems of Delphinium barbeyi (Ranunculaceae), 12 Sep. 2004, A.<br />

Ramaley, <strong>CBS</strong>-H 19868 (teleomorph), single ascospore isolates, <strong>CBS</strong> 121622 =<br />

CPC 11600, CPC 11601–11604.<br />

Substrate <strong>and</strong> distribution: On fading <strong>and</strong> decaying plant material,<br />

on living leaves (phylloplane fungus), as secondary invader, as an<br />

endophyte, isolated from air, soil, foodstuffs, paints, textiles <strong>and</strong><br />

numerous other materials; cosmopolitan.<br />

124


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 17. <strong>Cladosporium</strong> herbarum (CPC 11600). Macro- <strong>and</strong> micronematous conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

Literature: de Vries (1952: 71), Hughes (1958: 750), Ellis (1971:<br />

313), Domsch et al. (1980: 204), Sivanesan (1984: 225), Ellis &<br />

Ellis (1985: 290, 468, 1988: 168), Prasil & de Hoog (1988), Wang<br />

& Zabel (1990: 202), McKemy & Morgan-Jones (1991), Dugan &<br />

Roberts (1994), David (1997: 59), Ho et al. (1999: 129), de Hoog et<br />

al. (2000: 587), Samson et al. (2000: 110), Samson et al. (2001).<br />

Notes: De Vries (1952) incorrectly selected a specimen of Link’s<br />

herbarium at herb. B as lectotype. Prasil & de Hoog (1988) discussed<br />

this typification <strong>and</strong> designated one of Persoon’s original specimens<br />

as lectotype in which C. herbarum could be recognised. <strong>The</strong> latter<br />

material, which is in poor condition, could be re-examined within<br />

the course of these investigations <strong>and</strong> showed conidia agreeing<br />

with the current species concept of C. herbarum being (6–)9.5–<br />

14.5(–21) × (5–)6–7(–8) µm. Since the identity of the strain <strong>CBS</strong><br />

177.71 chosen by Prasil & de Hoog (1988) as representative living<br />

strain of C. herbarum could not be corroborated, an epitype with<br />

a living ex-epitype culture is designated. <strong>The</strong> holotype specimen<br />

of D. tassiana (RO) is morphologically <strong>similar</strong> to that observed on<br />

the epitype of C. herbarum, having ascospores which are (17–)21–<br />

www.studiesinmycology.org<br />

23(–25) × (6–)7–8(–8.5) µm, turning brown <strong>and</strong> verruculose in asci<br />

with age. However, no hamathecial remnants were observed in<br />

ascomata in vivo.<br />

<strong>The</strong> connection to the teleomorph D. tassiana could be<br />

confirmed, which is in agreement with the findings of von Arx (1950)<br />

<strong>and</strong> Barr (1958). Ascospore isolates formed the typical C. herbarum<br />

anamorph in culture, <strong>and</strong> these anamorph cultures developed<br />

some immature fruiting bodies within the agar. When inoculated<br />

onto water agar plates with nettle stems, numerous ascomata with<br />

viable ascospores were formed in culture.<br />

<strong>Cladosporium</strong> iridis (Fautrey & Roum.) G.A. de Vries, Contr.<br />

Knowl. Genus <strong>Cladosporium</strong>: 49. 1952. Figs 20–21.<br />

Basionym: Scolicotrichum iridis Fautrey & Roum., Rev. Mycol.<br />

(Toulouse) 13: 82. 1891.<br />

≡ Heterosporium iridis (Fautrey & Roum.) J.E. Jacques, Contr. Inst. Bot.<br />

Univ. Montréal 39: 18. 1941.<br />

For additional synonyms see Dugan et al. (2004).<br />

Teleomorph: Davidiella macrospora (Kleb.) Crous & U. Braun,<br />

Mycol. Progr. 2: 10. 2003.<br />

125


Schubert et al.<br />

Fig. 18. <strong>Cladosporium</strong> herbarum (CPC 11600) <strong>and</strong> its teleomorph Davidiella tassiana (from the host <strong>and</strong> CPC 12181). A–B. Macronematous conidiophores. C. Micronematous<br />

conidiophore. D. Microcyclic conidiogenesis. E. Conidial chain. F. Ascomata on the leaf. G. Ascomata formed in culture on nettle stems. H–I. Asci on the host. J–K. Ascospores<br />

in culture. L. Asci in culture. Scale bars: A, E, H, J–L = 10 µm, F–G, I = 200 µm.<br />

Basionym: Didymellina macrospora Kleb., Ber. Deutsch. Bot. Ges.<br />

42: 60, 1924. 1925.<br />

≡ Mycosphaerella macrospora (Kleb.) Jørst., Meld. Stat. Plantepatol. Inst.<br />

1: 20. 1945.<br />

Mycelium branched, 2–8 µm wide, septate, not constricted at the<br />

septa, hyaline to pale brown, smooth, walls slightly thickened,<br />

sometimes guttulate. Conidiophores very long, usually terminally<br />

arising from ascending hyphae, erect to subdecumbent, slightly<br />

to distinctly flexuous, geniculate-sinuous, usually several times,<br />

subnodulose due to geniculate, sympodial proliferation forming<br />

swollen lateral shoulders, unbranched, rarely branched, up to<br />

720 µm long, 6–11 µm wide, swellings 8–11(–14) µm wide,<br />

pluriseptate, often very regularly septate, not constricted at<br />

the septa, pale to medium olivaceous-brown, somewhat paler<br />

towards the apex, smooth to minutely verruculose, walls only<br />

slightly thickened. Conidiogenous cells integrated, terminal as<br />

well as intercalary, cylindrical-oblong, 15–55 µm long, proliferation<br />

percurrent to sympodial, usually with a single geniculation forming<br />

laterally swollen shoulders often below a septum, conidiogenous<br />

loci confined to swellings, usually one locus per swelling, rarely<br />

two, protuberant, (2–)2.5–4 µm diam, somewhat thickened<br />

<strong>and</strong> darkened-refractive. Conidia solitary, sometimes in short,<br />

unbranched chains, straight to curved, young conidia pyriform to<br />

subcylindrical, connection between conidiophore <strong>and</strong> conidium<br />

being rather broad, subhyaline to pale olivaceous-brown, walls<br />

slightly thickened, then enlarging <strong>and</strong> becoming more thick-walled,<br />

cylindrical-oblong, soleiform with age, both ends rounded, usually<br />

with a slightly to distinctly bulbous base, visible from a very early<br />

stage, but broadest part often towards the apex not at the base, (18–)<br />

30–75(–87) × (7–)10–16(–18) µm [av. ± SD, 53.3 (± 17.8) × 12.6 (±<br />

2.2) µm], (0–)2–6(–7)-septate, usually not constricted at the septa,<br />

rarely slightly constricted, septa often becoming sinuous with age,<br />

pale to medium olivaceous-brown, sometimes darker, verrucose to<br />

echinulate, walls thickened, especially in older conidia, up to 1 µm<br />

thick, hila protuberant, often stalk-like or conically prolonged, up<br />

to 2 µm long, (2–)2.5–3.5(–4) µm diam, with age becoming more<br />

sessile, sometimes just visible as a thickened plate just below the<br />

outer wall layer, especially in distal scars of branched conidia,<br />

periclinal rim often distinctly visible, hila somewhat thickened <strong>and</strong><br />

darkened-refractive; microcyclic conidiogenesis not observed.<br />

126


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 19. <strong>Cladosporium</strong> herbarum (CPC 11600). A. Overview of hyphal growth <strong>and</strong> conidiophore formation of a colony on SNA. Conidiophores are often formed on very wide<br />

(approx. 10 μm), septate hyphae that often grow near the agar surface. B. A more detailed view on colony organisation reveals the ornamented conidia. Note the septum near the<br />

conidiophore (arrow). C. Detail of spore ornamentation <strong>and</strong> hila on a secondary ramoconidium (arrow). Ornamentation is visible during early stages of spore formation (arrow).<br />

D. Structure of the conidiophore, illustrating the complex morphology of the spore-forming apparatus. In addition, secondary ramoconidia, conidia, <strong>and</strong> a hilum on the conidium<br />

are visible. E. Complex structure of the spore-forming apparatus. F. Details of secondary ramoconidia with complex scar-pattern on the right cell. G. Details of a secondary<br />

ramoconidium giving rise to conidia. Note the lack of ornamentation at the location of spore formation. Scale bars: A = 50 µm, B, F = 10 µm, C–E, G = 5 µm.<br />

Cultural characteristics: Colonies on PDA reaching 19–23 mm<br />

diam after 14 d at 25 ºC, pale greenish olivaceous, smoke-grey<br />

to olivaceous-grey due to abundant aerial mycelium, greenish<br />

olivaceous to olivaceous reverse, margin broad, regular, entire<br />

edge to slightly undulate, feathery, aerial mycelium abundantly<br />

formed, felty, fluffy, covering large parts of the colony, mainly in<br />

the central parts, high, growth low convex with a somewhat raised<br />

colony centre. Colonies on MEA reaching 9–23 mm diam after 14 d<br />

at 25 ºC, pale olivaceous-grey to olivaceous-grey, olivaceous-grey<br />

reverse, felty, margin slightly undulate, white, somewhat raised,<br />

aerial mycelium abundant, loose, diffuse, high, growth low convex,<br />

radially furrowed, slightly folded. Colonies on OA reaching 10–19<br />

mm diam after 14 d at 25 ºC, olivaceous, margin broad, undulate,<br />

white, aerial mycelium white, very high, loose, diffuse, hairy, growth<br />

flat, due to the mycelium low convex, without prominent exudates<br />

<strong>and</strong> sporulating on all media.<br />

Specimens examined: Isolated from Iris sp. (Iridaceae), <strong>CBS</strong> 107.20. France,<br />

Cote d’Or, Jardin de Noidan, on leaves of Iris germanica, Jul. 1880, F. Fautrey,<br />

Roumeguère, Fungi Sel. Gall. Exs. No. 5689, PC, lectotype of C. iridis, selected<br />

by David, 1997; K, isolectotype. Netherl<strong>and</strong>s, Boterenbrood, isolated from leaves<br />

of Iris sp., Aug. 1940, <strong>CBS</strong>-H 19859, epitype designated here of C. iridis, culture<br />

ex-epitype <strong>CBS</strong> 138.40.<br />

www.studiesinmycology.org<br />

127


Schubert et al.<br />

Fig. 20. <strong>Cladosporium</strong> iridis (<strong>CBS</strong> 138.40). Conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

Fig. 21. <strong>Cladosporium</strong> iridis (teleomorph Davidiella macrospora) (<strong>CBS</strong> 138.40). A–C. Conidiophores with conidia. D. Conidium. Scale bar = 10 µm.<br />

128


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 22. <strong>Cladosporium</strong> macrocarpum (<strong>CBS</strong> 299.67). Macro- <strong>and</strong> micronematous conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

Substrates <strong>and</strong> distribution: Leaf spot <strong>and</strong> blotch of Iris spp.<br />

including I. crocea, I. florentina, I. foetidissima, I. germanica, I.<br />

gueldenstaedtiana, I. kamaonensis, I. pallida, I. plicata (= I. swertii<br />

Hort.), I. pseudacorus, I. pumila, I. spuria ssp. halophila, <strong>and</strong> other<br />

species, also on Belacam<strong>and</strong>a chinensis (= Gemmingia chinensis),<br />

Hemerocallis fulva, Gladiolus g<strong>and</strong>avensis; Africa (Algeria,<br />

Morocco, South Africa, Zambia, Zimbabwe), Asia (Armenia,<br />

Azerbaijan, China, Georgia, India, Iran, Israel, Japan, Kazakhstan,<br />

Kirgizstan, Korea, Russia, Turkey, Turkmenistan, Uzbekistan),<br />

Australasia (Australia, New Zeal<strong>and</strong>), Europe (Austria, Belgium,<br />

Belorussia, Cyprus, Czech Republic, Denmark, Estonia, France,<br />

Germany, Great Britain, Greece, Italy, Latvia, Lithuania, Malta,<br />

Moldavia, Montenegro, Netherl<strong>and</strong>s, Norway, Pol<strong>and</strong>, Romania,<br />

Russia, Serbia, Spain, Sweden, Ukraine), North America (Canada,<br />

U.S.A.), Central & South America (Argentina, Chile, Jamaica,<br />

Panama, Uruguay).<br />

Literature: Ellis (1971: 312), Ellis & Waller (1974), Sivanesan (1984:<br />

222), McKemy & Morgan-Jones (1990), David (1997: 43), Shin et<br />

al. (1999).<br />

Notes: <strong>The</strong> description of the morphological parameters in culture<br />

is based on the isolate sporulating on PDA, since sporulation on<br />

SNA was not observed. <strong>The</strong> conidiophores <strong>and</strong> conidia in vivo are<br />

usually wider than in culture [conidiophores (6–)9–15(–17) µm<br />

wide, conidia (11–)15–23(–28) µm].<br />

www.studiesinmycology.org<br />

<strong>Cladosporium</strong> macrocarpum Preuss, in Sturm, Deutsch. Fl.<br />

3(26): 27. 1848. Figs 22–25.<br />

≡ <strong>Cladosporium</strong> herbarum var. macrocarpum (Preuss) M.H.M. Ho &<br />

Dugan, in Ho et al., Mycotaxon 72: 131. 1999.<br />

= Dematium herbarum var. (β) brassicae Pers., Syn. meth. fung. 2: 699. 1801,<br />

syn. nov.<br />

= Dematium graminum Pers., Mycol. eur. 1: 16. 1822, syn. nov.<br />

= Dematium vulgare var. (δ) typharum Pers., Mycol. eur. 1: 14. 1822, syn.<br />

nov.<br />

= Dematium vulgare var. (β) foliorum Pers., Mycol. eur. 1: 14. 1822, syn. nov.<br />

For additional synonyms see Dugan et al. (2004), Schubert<br />

(2005).<br />

Teleomorph: Davidiella macrocarpa Crous, K. Schub. & U. Braun,<br />

sp. nov. MycoBank MB504582.<br />

Davidiellae tassianae similis, sed pseudoparaphysibus prominentibus et ascosporis<br />

maioribus, (22–)23–26(–28) × (6–)6.5–7(–8) µm.<br />

Ascomata superficial on a small stroma, black, up to 200 µm<br />

diam, globose, separate, but developing with 1–3 necks with age;<br />

ostioles consisting of pale brown to subhyaline cells, periphysate,<br />

with periphysoids growing into the cavity; wall consisting of 3–6<br />

layers of medium brown textura angularis. Pseudoparaphyses<br />

present, hyaline, subcylindrical, septate, anastomosing, 3–4<br />

µm diam; hamathecial cells persistent in cavity. Asci fasciculate,<br />

bitunicate, subsessile, broadly ellipsoid with a long tapered stalk,<br />

straight to curved, 8-spored, 70–110 × 15–20 µm. Ascospores<br />

tri- to multiseriate, overlapping, hyaline, guttulate, irregular lumina<br />

129


Schubert et al.<br />

Fig. 23. <strong>Cladosporium</strong> macrocarpum (<strong>CBS</strong> 299.67) <strong>and</strong> its teleomorph Davidiella macrocarpa (CPC 12755). A–C. Macronematous conidiophores <strong>and</strong> conidia. D–G.<br />

Micronematous conidiophores. H. Microcyclic conidiogenesis. I. Ascomata formed on nettle stems in culture. J. Periphyses. K, M–N. Asci. L. Ostiole. Scale bars: A, D–H, J–N<br />

= 10 µm, I = 200 µm.<br />

rarely observed, thick-walled, straight to slightly curved, fusoidellipsoidal<br />

with obtuse ends, widest in the middle of the apical<br />

cell, medianly 1-septate, not to slightly constricted at the septum,<br />

tapering towards both ends, but more prominently towards lower<br />

end, (22–)23–26(–28) × (6–)6.5–7(–8) µm; mucoid sheath rarely<br />

observed, mostly absent.<br />

Mycelium unbranched or loosely branched, 1–4.5(–5) µm wide,<br />

septate, sometimes slightly constricted at septa, hyaline to pale<br />

brown, smooth to minutely verruculose, walls unthickened or slightly<br />

thickened. Conidiophores micronematous <strong>and</strong> macronematous,<br />

solitary, arising terminally from plagiotropous hyphae or terminally<br />

from ascending hyphae. Macronematous conidiophores erect,<br />

straight to somewhat flexuous, cylindrical-oblong, nodulose to<br />

nodose, with a single apical or usually several swellings either<br />

somewhat distinct from each other or often in short succession giving<br />

conidiophores a knotty appearance, swellings sometimes laterally<br />

elongated or formed at the top of a branch-like outgrowth below<br />

the apical swelling, sometimes distinctly geniculate, unbranched,<br />

sometimes branched, 12–260 × (3–)4–6 µm, swellings 5–10 µm<br />

wide, pluriseptate, sometimes slightly constricted at septa, pale to<br />

medium brown or olivaceous-brown, somewhat paler at apices,<br />

130


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 24. <strong>Cladosporium</strong> macrocarpum (<strong>CBS</strong> 299.67). A. Survey of a conidiophore that forms several secondary ramoconidia <strong>and</strong> conidia. Several aerial hyphae are also visible<br />

in this picture. B. Conidiophore with broadly ellipsoid secondary ramoconidia <strong>and</strong> obovoid conidia. Note the different scars on the conidiophore at the lower left. C. Ellipsoid<br />

or obovoid conidia with notable areas of scar formation. <strong>The</strong> ornamentation is relatively widely distributed over the body of the cell <strong>and</strong> <strong>similar</strong> to C. variabile. D. Detail of<br />

a conidiophore (see B) with scars. Note the relatively shallow rings of the scars. E. Details of conidia <strong>and</strong> a secondary ramoconidium. F. Conidiophore with a secondary<br />

ramoconidium <strong>and</strong> conidia. Note the hila on several spores <strong>and</strong> the lack of ornamentation at the site where spores are formed. Scale bars: A–C, = 10 µm, D, F = 5 µm, E = 2<br />

µm.<br />

smooth to minutely verruculose or verruculose, walls somewhat<br />

thickened, sometimes even two-layered. Conidiogenous cells<br />

integrated, terminal or intercalary, cylindrical, nodulose with lateral<br />

shoulders or nodose with swellings round about the stalk, with<br />

conidiogenous loci confined to swellings, 12–37 µm long, with up to<br />

12 loci per cell, usually with up to six, loci conspicuous, protuberant,<br />

(1–)1.5–2 µm diam, somewhat thickened <strong>and</strong> darkened-refractive.<br />

Micronematous conidiophores almost indistinguishable from<br />

hyphae, straight, narrowly filiform, non-nodulose or with a single or<br />

few swellings, mostly with small head-like swollen apices, usually<br />

www.studiesinmycology.org<br />

131


Schubert et al.<br />

or branched chains, subglobose, obovoid to limoniform, ellipsoid<br />

or fusiform, 2.5–16 × 1.5–5 µm, 0(–1)-septate, few longer conidia<br />

subcylindrical to clavate, up to 37(–43) µm long, 0–2(–3)-septate,<br />

occasionally with up to four septa, sometimes slightly constricted<br />

at the septa, subhyaline to pale brown, almost smooth to minutely<br />

verruculose, walls unthickened, hila 0.8–1.2 µm diam, thickened<br />

<strong>and</strong> darkened-refractive.<br />

Cultural characteristics: Colonies on PDA reaching 30–43 mm<br />

in diam after 14 d at 25 ºC, dark dull green to olivaceous-grey,<br />

olivaceous-grey, dark olivaceous- to iron-grey reverse, pulvinate,<br />

velvety, sometimes somewhat zonate, paler zones towards the<br />

margin, margin regular, entire edge, almost colourless to white,<br />

glabrous to feathery, aerial mycelium sparse to more abundant<br />

in the colony centre or covering large areas of the colony, hairy,<br />

fluffy or felty, whitish to smoke-grey, sometimes becoming reddish,<br />

livid red to vinaceous, growth flat, regular, sometimes forming<br />

few prominent exudates, exudates sometimes slightly reddish,<br />

sporulation profuse with two kinds of conidiophores, low <strong>and</strong> high.<br />

Colonies on MEA reaching 31–50 mm in diam after 14 d at 25<br />

ºC, grey-olivaceous to olivaceous-grey or iron-grey, sometimes<br />

pale olivaceous-grey to whitish due to abundant aerial mycelium,<br />

olivaceous-grey or iron-grey reverse, velvety or powdery, margin<br />

narrow, entire edge, colourless to white, glabrous, aerial mycelium<br />

sparse to abundant, hairy or felty, growth regular, flat to low convex,<br />

radially furrowed, without prominent exudates, sporulation profuse.<br />

Colonies on OA reaching 29–40 mm in diam after 14 d at 25 ºC,<br />

grey-olivaceous, olivaceous-grey to dark smoke-grey, olivaceousblack<br />

or iron grey reverse, margin entire edge, narrow, colourless<br />

or white, glabrous, aerial mycelium sparse, mainly in the colony<br />

centre, felty, white to smoke-grey or grey-olivaceous, felty, growth<br />

flat, regular, without exudates, sporulating.<br />

Fig. 25. Davidiella macrocarpa (CPC 12755). Ascus <strong>and</strong> ascospores. Scale bar =<br />

10 µm. P.W. Crous del.<br />

only few micrometer long, 1.5–3 µm wide, aseptate or with only<br />

few septa, subhyaline, smooth or almost so, walls unthickened,<br />

with a single or only few conidiogenous loci, narrow, 0.8–1.2 µm<br />

diam, thickened <strong>and</strong> somewhat darkened-refractive. Conidia<br />

catenate, in branched chains, small terminal conidia subglobose,<br />

obovoid, oval, limoniform, 4–11 × (3–)4–6 µm [av. ± SD, 7.6 (±<br />

1.9) × 5.0 (± 0.8) µm], aseptate, intercalary conidia broadly ovoidellipsoid,<br />

10–17 × (4.5–)5–9 µm [av. ± SD, 12.7 (± 2.1) × 6.8 (±<br />

0.8) µm], 0–1-septate; secondary ramoconidia broadly ellipsoid<br />

to subcylindrical, 14–25(–30) × (5–)6–9(–10) µm [av. ± SD, 19.4<br />

(± 3.5) × 7.6 (± 1.0) µm], 0–2(–3)-septate, sometimes slightly<br />

constricted at the septa, septa somewhat sinuous with age, pale<br />

brown to medium olivaceous-brown or brown, sometimes even<br />

dark brown, verruculose to echinulate (muricate under SEM),<br />

walls thickened, up to 1 µm thick, mostly broadly rounded at<br />

apex <strong>and</strong> base, sometimes attenuated, sometimes guttulate<br />

by oil drops, with up to three apical hila, mostly 1–2, hila sessile<br />

(apparently somewhat immersed) to somewhat protuberant, 1–<br />

2(–2.5) µm diam, thickened <strong>and</strong> darkened-refractive; microcyclic<br />

conidiogenesis occurring with conidia forming secondary micro<strong>and</strong><br />

macronematous conidiophores, conidia often germinating with<br />

long hyphae. Conidia formed by micronematous conidiophores<br />

usually smaller, narrower <strong>and</strong> paler, catenate, in short unbranched<br />

Specimens examined: Sine loco et dato, L 910.255-723 = L-0115836, lectotype<br />

designated here of Dematium graminum. Sine loco, on dead stems of Brassica sp.<br />

(Brassicaceae), No. 601, L 910.255-716 = L-0115849, holotype of D. herbarum var.<br />

(β) brassicae. Sine loco, on leaves of Iris (Iridaceae), Quercus (Fagaceae), Brassica<br />

etc., L 910.255-736 = L-0115871, holotype of D. vulgare var. (β) foliorum, isotype L<br />

910.255-718 = L-0115872. Sine loco et dato, L 910.255-698 = L-0115852, lectotype<br />

designated here for D. vulgare var. (δ) typharum. Isolated from “Mycosphaerella<br />

tulasnei”, <strong>CBS</strong> 223.32 = ATCC 11287 = IMI 049635. Romania, isolated from water,<br />

<strong>CBS</strong> 175.82. Slovenia, Sečovlje, isolated from hypersaline water from salterns<br />

(precrystalisation pond), 2004, P. Zalar, EXF-2287 = CPC 12054. Turkey, Ankara,<br />

Tekeli, isolated from Triticum aestivum (Poaceae), isol. S. Tahsin, ident. A.C. Stolk,<br />

<strong>CBS</strong> 299.67. U.S.A., Seattle, University of Washington Campus, 47.6263530, -<br />

122.3331440, isolated from cleistothecia of Phyllactinia guttata (Erysiphaceae)<br />

on leaves of Corylus sp. (Corylaceae), 16 Sep. 2004, D. Glawe, CPC 11817;<br />

Washington, isolated from Spinacia oleracea (Chenopodiaceae), 1 Jan. 2003,<br />

L. DuToit, <strong>CBS</strong>-H 19855, neotype designated here for C. macrocarpum, <strong>and</strong><br />

holotype of D. macrocarpa, isoneotype HAL 2020 F, isotype HAL 2021 F, culture<br />

ex-type CPC 12752, 12756–12759, CPC 12755 = <strong>CBS</strong> 121623.<br />

Substrate <strong>and</strong> distribution: Decaying plant material, human,<br />

hypersaline water, water; widespread.<br />

Literature: de Vries (1952: 76), Ellis (1971: 315), Domsch et al.<br />

(1980: 208), Ellis & Ellis (1985: 290, 468), Matsushima (1985: 5),<br />

McKemy & Morgan-Jones (1991), Dugan & Roberts (1994), David<br />

(1997: 71), Samson et al. (2000: 112).<br />

Notes: In the absence of Preuss’s type material (not preserved)<br />

de Vries (1952) “lectotypified” C. macrocarpum by a specimen in<br />

Saccardo’s herbarum (Herb. Myc. P.A. Saccardo no. 419, PAD).<br />

This material, subsequently distributed in Mycotheca Italica no.<br />

1396, should correctly be regarded as neotype (David 1997). A<br />

single collection of Saccardo’s Mycotheca Italica no. 1396 from<br />

herb. HBG, which can be considered as isoneotype material,<br />

132


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 26. <strong>Cladosporium</strong> ossifragi (<strong>CBS</strong> 842.91). Conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

was re-examined <strong>and</strong> proved to rather agree with the species<br />

concept of C. herbarum s. str. <strong>The</strong> conidia were formed in simple,<br />

rarely branched chains, 6–26 × (4–)5.5–8(–9) µm, 0–3-septate,<br />

almost smooth or minutely to densely verruculose or verrucose<br />

(Schubert 2005). However, since de Vries’ “lectotypification” was<br />

incorrect according to the code (ICBN, Art. 9.2, 9.17), a neotype is<br />

designated.<br />

<strong>The</strong> delimitation of C. macrocarpum as a morphologically<br />

distinct species from C. herbarum has been controversially<br />

discussed by several authors (McKemy & Morgan-Jones 1991,<br />

Dugan & Robert 1994, Ho et al. 1999). Based on molecular as well<br />

morphological studies, it can be shown that C. macrocarpum is a<br />

well-defined species distinguishable from C. herbarum s. str. by<br />

forming conidiophores with wider nodes, 5–10 µm, wider <strong>and</strong> more<br />

frequently septate conidia [small terminal conidia 4–11 × (3–)4–6<br />

µm versus 4–10 × 3–5(–6) µm in C. herbarum, intercalary conidia<br />

10–17 × (4.5–)5–9 µm versus 6–16 × 4–6 µm in C. herbarum,<br />

secondary ramoconidia 14–25(–30) × (5–)6–9(–10) µm versus 12–<br />

25(–35) × (3–)5–7(–9) µm in C. herbarum] <strong>and</strong> by being connected<br />

to Davidiella macrocarpa. On natural substrates the conidiophores<br />

are usually somewhat wider than in culture, 4–8(–10) µm wide, <strong>and</strong><br />

also the conidia can be somewhat wider, sometimes up to 13(–15) µm.<br />

<strong>Cladosporium</strong> graminum, described by Persoon (1822), as<br />

well as C. brunneum <strong>and</strong> C. gracile, introduced by Corda (1837),<br />

are older synonyms of C. macrocarpum <strong>and</strong>, according to the<br />

code, would have priority. However, since C. macrocarpum is a<br />

well established, currently used name with numerous records in<br />

literature, a proposal to conserve the name against these older<br />

names is in preparation for formal publication in Taxon.<br />

A characteristic difference between ascomata of C.<br />

macrocarpum in comparison to those of C. herbarum, are the<br />

smaller, globose pseudothecia, asci with longer stalks, prominence<br />

of pseudoparaphyses, <strong>and</strong> rather inconspicuous luminar ascospore<br />

inclusions.<br />

<strong>Cladosporium</strong> ossifragi (Rostr.) U. Braun & K. Schub., comb.<br />

nov. MycoBank MB504575. Figs 26–28.<br />

Basionym: Napicladium ossifragi Rostr., Bot. Fǽröes 1: 316.<br />

1901.<br />

≡ Heterosporium ossifragi (Rostr.) Lind, Dan. fung.: 531. 1913.<br />

= Heterosporium magnusianum Jaap, Schriften Naturwiss. Vereins Schleswig-<br />

Holstein 12: 346. 1902.<br />

≡ <strong>Cladosporium</strong> magnusianum (Jaap) M.B. Ellis in Ellis, More<br />

Dematiaceous Hyphomycetes: 337. 1976.<br />

www.studiesinmycology.org<br />

133


Schubert et al.<br />

Fig. 27. <strong>Cladosporium</strong> ossifragi (<strong>CBS</strong> 842.91). A. Macronematous conidiophore. B. Micronematous conidiophore. C–D. Conidia. E. Conidia <strong>and</strong> microcyclic conidiogenesis.<br />

Scale bars = 10 µm.<br />

Fig. 28. <strong>Cladosporium</strong> ossifragi (<strong>CBS</strong> 842.91). A. Survey on different secondary ramoconidia <strong>and</strong> conidia. B. Details of conidia <strong>and</strong> hila. Note the very pronounced ornamentation<br />

<strong>and</strong> the absence of ornamentation near the site of spore formation. C. Detail of the end of a secondary ramoconidium with pronounced hila. D. Formation of a new conidium.<br />

Note the broad scar behind it (> 1 μm). E. Formation of a new conidium from a smooth-walled stalk. F. Hila on a secondary ramoconidium. This micrograph is from the sample<br />

before coating with gold-palladium <strong>and</strong> shows <strong>similar</strong> features as the sample after sputter coating. Scale bars: A = 10 µm, B–D, F = 2 µm, E = 5 µm.<br />

134


<strong>Cladosporium</strong> herbarum species complex<br />

Mycelium abundantly formed, twisted, often somewhat aggregated,<br />

forming ropes, branched, 1–5 µm wide, septate, often irregularly<br />

swollen <strong>and</strong> constricted, hyaline or subhyaline to pale brown,<br />

smooth, walls unthickened or only slightly thickened. Conidiophores<br />

macronematous <strong>and</strong> micronematous, arising from plagiotropous<br />

hyphae, terminally or laterally, erect to subdecumbent, more<br />

or less straight to flexuous, cylindrical, sometimes geniculate,<br />

subnodulose with loci often situated on small lateral shoulders,<br />

unbranched, sometimes branched, often very long, up to 350 µm<br />

long, 3–4.5(–5) µm wide, pluriseptate, shorter ones aseptate, not<br />

constricted at septa, pale to pale medium brown, paler towards<br />

apices, sometimes subhyaline, smooth to minutely verruculose,<br />

especially towards apices, walls somewhat thickened, up to 0.5 µm,<br />

sometimes appearing two-layered. Conidiogenous cells integrated,<br />

terminal as well as intercalary, cylindrical, sometimes geniculate,<br />

subnodulose, 5–31 µm long, proliferation sympodial, with few<br />

loci (1–3) per cell, loci usually confined to small lateral shoulders,<br />

protuberant, conspicuous, short cylindrical, 1–2 µm wide, up to<br />

1 µm high, somewhat thickened, darkened-refractive. Conidia<br />

catenate, in short, unbranched or branched chains, straight, small<br />

terminal <strong>and</strong> intercalary conidia subglobose, obovoid to ellipsoid,<br />

4–15 × 3–5 µm [av. ± SD, 9.3 (± 3.7) × 4.0 (± 0.7) µm], 0–1-<br />

septate, not constricted at the septa, pale brown, hila 0.8–1 µm<br />

diam, secondary ramoconidia cylindrical, sometimes ellipsoid or<br />

subfusiform, 16–36(–40) × (4–)5–8 µm [av. ± SD, 26.6 (± 7.4) ×<br />

6.0 (± 1.2) µm], (0–)1–3(–4)-septate [in vivo wider, (6–)7–9(–11)<br />

µm, <strong>and</strong> with up to five, rarely seven septa], not constricted at<br />

the septa, septa sometimes slightly sinuous, pale brown to pale<br />

medium brown, densely verruculose, verrucose to echinulate<br />

(densely muricate under SEM), walls unthickened to somewhat<br />

thickened, rounded or somewhat attenuated at apex <strong>and</strong> base,<br />

hila protuberant, conspicuous, sometimes situated on short, small<br />

prolongations, 1–2.5 µm diam, somewhat thickened <strong>and</strong> darkenedrefractive;<br />

microcyclic conidiogenesis occasionally occurring.<br />

Cultural characteristics: Colonies on PDA reaching 53 mm diam<br />

after 14 d at 25 ºC, greenish olivaceous, grey-olivaceous to<br />

olivaceous-grey or iron-grey, appearing somewhat zonate, dull<br />

green to olivaceous-black reverse, margin colourless, regular,<br />

entire edge, aerial mycelium abundantly formed, covering at first<br />

the colony centre later most of the surface, dense, high, growth flat<br />

with elevated colony centre, somewhat folded. Colonies on MEA<br />

reaching 54 mm diam after 14 d at 25 ºC, pale olivaceous-grey<br />

to olivaceous-grey in the centre, iron-grey reverse, velvety, margin<br />

colourless to white, entire edge, radially furrowed, aerial mycelium<br />

abundantly formed, fluffy to felty, growth flat with somewhat raised,<br />

folded colony centre. Colonies on OA attaining 52 mm diam after<br />

14 d at 25 ºC, olivaceous-grey to iron-grey, iron-grey to greenish<br />

black reverse, margin white, entire edge, aerial mycelium diffuse,<br />

loose, growth flat, prominent exudates absent, sporulation profuse<br />

on all media.<br />

Specimens examined: Denmark, Undallslund, on leaves of Narthecium ossifragum<br />

(Melanthiaceae), 13 Sep. 1885, E. Rostrup, CP, neotype designated here of C.<br />

ossifragi; Tǿnder, Rǿmǿ near Twismark, 19 Aug. 1911, H. Sydow, Sydow, Mycoth.<br />

Germ. 1047, M. Germany, Hamburg, Eppendorfer Moor, on leaves of Narthecium<br />

ossifragum, 12 Sep. 1897, O. Jaap, HBG, lectotype selected here of C.<br />

magnusianum; 4 Sep. 1903, O. Jaap, Jaap, Fungi Sel. Exs. 49, M; Wernerwald near<br />

Cuxhaven, Aug. 1927, A. Ludwig, Petrak, Mycoth. Gen. 146, M. Norway, Bjerkreim<br />

County, isolated from leaves of Narthecium ossifragum, M. di Menna, <strong>CBS</strong>-H 19860,<br />

epitype designated here of C. ossifragi, culture ex-epitype <strong>CBS</strong> 842.91 = ATCC<br />

200946; Møre og Romsdal County, isolated from leaves of Narthecium ossifragum,<br />

M. di Menna, <strong>CBS</strong> 843.91.<br />

Substrate <strong>and</strong> distribution: Causing leaf spots on Narthecium<br />

ossifragum; Europe (Austria, Denmark, Germany, Great Britain,<br />

Irel<strong>and</strong>, Norway).<br />

Literature: Ellis & Ellis (1985: 390), David (1995a; 1997: 85–86,<br />

88), Ho et al. (1999: 132).<br />

Notes: Type material of Napicladium ossifragi is not preserved in<br />

Rostrup’s herbarium (on Narthecium ossifragum, Faeroe Isl<strong>and</strong>s,<br />

Viderö, Viderejde <strong>and</strong> Österö, Svinaa, sine dato, leg. Ostenfeld &<br />

Harz). However, other authentic collections seen <strong>and</strong> examined<br />

by Rostrup are deposited at CP. Lind (1913) re-examined these<br />

samples, synonymised N. ossifragi with H. magnusianum <strong>and</strong><br />

correctly introduced the combination H. ossifragi. Nevertheless,<br />

the correct oldest name for this fungus has been ignored by<br />

most authors. David (1997), who clearly stated that N. ossifragi is<br />

the earliest name for this species, preferred to use the name C.<br />

magnusianum because the typification of Rostrup’s name was still<br />

uncertain. Despite the lacking type material, there is no doubt about<br />

the correct identity of N. ossifragi since authentic material of this<br />

species, examined by <strong>and</strong> deposited in Rostrup’s herbarium (CP),<br />

is preserved. <strong>The</strong>refore, there is no reason to reject the oldest valid<br />

name for this species. <strong>The</strong> original collection of C. magnusianum<br />

cited by Jaap (1902) (on leaves of Narthecium ossifragum, Denmark,<br />

Tǿnder, Rǿmǿ, peatbog by Twismark, Jul.–Aug. 1901, Jaap), but<br />

not designated as type, is not preserved (David 1997). It is neither<br />

deposited at B, HBG nor S. However, in the protologue Jaap (1902)<br />

also referred to material of this species found near Hamburg, which<br />

is, hence, syntype material available for lectotypification.<br />

<strong>Cladosporium</strong> pseudiridis K. Schub., C.F. Hill, Crous & U. Braun,<br />

sp. nov. MycoBank MB504576. Figs 29–30.<br />

Etymology: Epithet derived from its <strong>similar</strong> morphology to<br />

<strong>Cladosporium</strong> iridis.<br />

Differt a Cladosporio iridis conidiis 0–3-septatis, brevioribus et latioribus, 15–55 ×<br />

(9–)11–19(–21) µm.<br />

Mycelium sparingly branched, 2–7 µm wide, septate, not constricted<br />

at the septa, subhyaline to pale brown, smooth or almost so, walls<br />

somewhat thickened, guttulate or protoplasm appearing granular,<br />

sometimes enveloped by a slime coat. Conidiophores arising<br />

mostly terminally from ascending hyphae, sometimes also laterally<br />

from plagiotropous hyphae, erect, more or less straight, broadly<br />

cylindrical-oblong, once or several times slightly to distinctly<br />

geniculate-sinuous, forming more or less pronounced lateral<br />

shoulders, nodulose, unbranched, 100–320(–500) × 7–11 µm,<br />

swellings 10–14 µm wide, becoming narrower <strong>and</strong> paler towards the<br />

apex, septate, not constricted at the septa, septa mainly basal, apical<br />

cell often very long, pale to medium olivaceous-brown, subhyaline<br />

at the apex, smooth or almost so, sometimes minutely verruculose,<br />

walls usually distinctly thickened, sometimes even two-layered, up<br />

to 1(–2) µm thick, protoplasm granular, often clearly contrasting<br />

from the outer wall. Conidiogenous cells integrated, terminal <strong>and</strong><br />

intercalary, cylindrical-oblong, slightly to distinctly geniculatesinuous,<br />

nodulose with conidiogenous loci confined to swellings<br />

or lateral shoulders, 30–110 µm long, proliferation percurrent to<br />

sympodial, with a single or three, sometimes up to five geniculations<br />

per cell, usually only a single locus per swelling, protuberant, very<br />

prominent, short cylindrical, peg-like, clearly composed of a dome<br />

<strong>and</strong> surrounding rim, dome often higher than the periclinal rim,<br />

broad, somewhat paler than rim, conically narrowed, (2–)2.5–4<br />

µm wide, up to 2 µm high, thickened <strong>and</strong> darkened-refractive.<br />

www.studiesinmycology.org<br />

135


Schubert et al.<br />

Fig. 29. <strong>Cladosporium</strong> pseudiridis (<strong>CBS</strong> 116463). Conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

Conidia solitary, sometimes in short unbranched chains of two or<br />

three, straight to slightly curved, young conidia small, 0–1-septate,<br />

broadly ovoid to pyriform, 15–26 × (9–)11–16(–18) µm [av. ± SD,<br />

19.2 (± 4.3) × 14.2 (± 3) µm], first septum somewhat in the upper<br />

half, the upper cell is much smaller but gradually extending as the<br />

conidium matures, mature conidia 1–3-septate, broadly pyriform,<br />

cylindrical-oblong or soleiform, usually with a distinctly bulbous<br />

base, 30–55 × 12–19(–21) µm [av. ± SD, 41.5 (± 6.8) × 17.1 (± 2.1)<br />

µm], broadest part of conidia usually at the bulbous base, mostly<br />

attenuated towards the basal septum, septa becoming sinuous with<br />

age, pale to medium olivaceous-brown or brown, usually echinulate,<br />

sometimes coarsely verrucose, walls distinctly thickened, up to 2<br />

µm thick, often appearing layered with a large lumen in the centre<br />

of the cell, broadly rounded to flattened at apex <strong>and</strong> base, hila often<br />

very prominent, often peg-like elongated, up to 3 µm long, with age<br />

becoming less prominent, visible as a thickened flat plate just below<br />

the outer echinulate wall layer, slightly raised towards the middle,<br />

2–3.5 µm diam, thickened <strong>and</strong> darkened-refractive; microcyclic<br />

conidiogenesis not observed.<br />

Cultural characteristics: Colonies on PDA attaining 6 mm diam<br />

after 14 d at 25 ºC, whitish, smoke-grey to pale olivaceous-grey<br />

due to abundant aerial mycelium, olivaceous-black reverse, margin<br />

narrow, white, more or less crenate, aerial mycelium zonate, fluffy,<br />

covering most of the colony, mainly in the colony centre, growth<br />

136


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 30. <strong>Cladosporium</strong> pseudiridis (<strong>CBS</strong> 116463). A–C. Conidiophores <strong>and</strong> conidia. D. Part of a conidiogenous cell showing a protuberant cladosporioid conidiogenous locus.<br />

E–F. Conidia. Scale bars = 10 µm.<br />

convex to raised, deep into the agar, with age few large prominent<br />

exudates formed, sparingly sporulating. Colonies on MEA attaining<br />

7 mm diam after 14 d at 25 ºC, olivaceous-grey, pale olivaceousgrey<br />

to pale rosy-buff due to abundant aerial mycelium covering<br />

almost the whole colony, iron-grey reverse, margin colourless or<br />

white, broad, regular, more or less glabrous, aerial mycelium fluffy,<br />

dense, high, growth convex to umbonate, sometimes with elevated<br />

colony centre, prominent exudates lacking, sporulation sparse.<br />

Colonies on OA attaining 8 mm diam after 14 d at 25 ºC, white, pale<br />

buff to pale olivaceous-grey in the centre, margin grey-olivaceous,<br />

olivaceous- to iron-grey reverse, margin entire edge or somewhat<br />

undulate, somewhat feathery, growth raised with a somewhat<br />

depressed centre forming an elevated outer rim, without prominent<br />

exudates, sporulation more abundant.<br />

Specimen examined: New Zeal<strong>and</strong>, Auckl<strong>and</strong>, Mt. Albert, Carrington Road, Unitec<br />

Campus, isolated from large leaf lesions on Iris sp. (Iridaceae), 15 Aug. 2004, C.F.<br />

Hill, <strong>CBS</strong>-H 19861, holotype, culture ex-type <strong>CBS</strong> 116463 = LYN 1065 = ICMP<br />

15579.<br />

Substrate <strong>and</strong> distribution: On living leaves of Iris sp.; New<br />

Zeal<strong>and</strong>.<br />

Notes: <strong>Cladosporium</strong> pseudiridis closely resembles C. iridis, a<br />

common <strong>and</strong> widespread species causing leaf spots on numerous<br />

Iris spp. <strong>and</strong> a few additional hosts of the host family Iridaceae,<br />

but the latter species is easily distinguishable by having longer<br />

<strong>and</strong> narrower, more frequently septate conidia, (18–)30–75(–87) ×<br />

(7–)10–16(–18) µm, (0–)2–6(–7)-septate.<br />

www.studiesinmycology.org<br />

It is unlikely that C. pseudiridis is of New Zeal<strong>and</strong> origin since<br />

the <strong>genus</strong> Iris is not indigenous to New Zeal<strong>and</strong>. All Iris species<br />

that are found in this country have been introduced, mainly for<br />

horticultural purposes. <strong>The</strong> species is, therefore, probably more<br />

common than indicated above. However, within the course of the<br />

recent monographic studies in the <strong>genus</strong> <strong>Cladosporium</strong> numerous<br />

herbarium specimens, mainly of European origin, have been<br />

examined <strong>and</strong> proved to be correctly identified agreeing with the<br />

species concept of C. iridis. Additional collections <strong>and</strong> cultures are<br />

necessary to determine its distribution.<br />

<strong>Cladosporium</strong> ramotenellum K. Schub., Zalar, Crous & U. Braun,<br />

sp. nov. MycoBank MB504577. Figs 31–33.<br />

Etymology: Refers to the morphological <strong>similar</strong>ity with <strong>Cladosporium</strong><br />

tenellum.<br />

Differt a Cladosporio cladosporioide conidiophoris et conidiis leniter angustioribus,<br />

2–4(–5) µm latis, conidiis 0–2(–3)-septatis, semper verruculosis; et a Cladosporio<br />

tenello locis conidiogenis non numerosis et non aggregatos ad apicem, conidiis<br />

longioribus et angustioribus, 2.5–35 × 2–4(–5) µm, 0–3-septatis.<br />

Mycelium unbranched or only sparingly branched, 1.5–4 µm wide,<br />

septate, without swellings <strong>and</strong> constrictions, hyaline or subhyaline,<br />

smooth, sometimes irregularly rough-walled, walls unthickened.<br />

Conidiophores solitary, macronematous <strong>and</strong> micronematous,<br />

arising as lateral branches of plagiotropous hyphae or terminally<br />

from ascending hyphae, erect, straight or slightly flexuous,<br />

cylindrical, neither geniculate nor nodulose, without head-like<br />

swollen apices or intercalary swellings, unbranched, sometimes<br />

137


Schubert et al.<br />

Fig. 31. <strong>Cladosporium</strong> ramotenellum (CPC 12043). Conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

branched, branches often only as short lateral prolongations,<br />

mainly formed below a septum, 14–110 × 2–4 µm, septate, not<br />

constricted at the septa, subhyaline to pale olivaceous or brown,<br />

smooth to minutely verruculose, walls unthickened, sometimes<br />

guttulate. Conidiogenous cells integrated, terminal, sometimes also<br />

intercalary, cylindrical, not geniculate, non-nodulose, 10–28(–50)<br />

µm long, proliferation sympodial, with few conidiogenous loci,<br />

mostly 1–3, loci sometimes situated on small lateral prolongations,<br />

protuberant, 0.5–1.5(–2) µm diam, thickened <strong>and</strong> somewhat<br />

darkened-refractive. Ramoconidia formed, cylindrical-oblong, up<br />

to 47 µm long, 2–4 µm wide, 0–1-septate, rarely up to 4-septate,<br />

subhyaline to very pale olivaceous, smooth or almost so, with a<br />

broadly truncate base, without any dome <strong>and</strong> raised rim, 2–3 µm<br />

wide, not thickened but somewhat refractive. Conidia numerous,<br />

polymorphous, catenate, in branched chains, straight, sometimes<br />

slightly curved, small terminal conidia numerous, globose,<br />

subglobose or ovoid, obovoid or limoniform, 2.5–7 × 2–4(–4.5) µm<br />

[av. ± SD, 5.1 (± 1.3) × 3.1 (± 0.6) µm], aseptate, without distal<br />

hilum or with a single apical scar, intercalary conidia ellipsoid to<br />

subcylindrical, 8–15 × 3–4(–4.5) µm [av. ± SD, 11.5 (± 2.4) × 3.6<br />

(± 0.5) µm], 0–1-septate; secondary ramoconidia subcylindrical<br />

to cylindrical-oblong, 17–35 × 3–4(–5) µm [av. ± SD, 22.5 (±<br />

5.6) × 3.7 (± 0.5) µm], 0–3-septate, not constricted at the septa,<br />

subhyaline to very pale olivaceous, minutely verruculose (granulate<br />

under SEM), walls unthickened or almost so, apex broadly rounded<br />

or slightly attenuated towards apex <strong>and</strong> base, sometimes guttulate,<br />

hila protuberant, conspicuous, 0.8–1.5(–2) µm diam, somewhat<br />

thickened <strong>and</strong> darkened-refractive; microcyclic conidiogenesis<br />

occurring.<br />

Cultural characteristics: Colonies on PDA reaching 46–49 mm diam<br />

after 14 d at 25 ºC, olivaceous to grey-olivaceous due to abundant<br />

sporulation, appearing zonate in forming concentric zones, margin<br />

entire edge to slightly undulate, white, glabrous, aerial mycelium<br />

absent or sparse, growth flat with a somewhat folded <strong>and</strong> wrinkled<br />

colony centre, without prominent exudates, sporulation profuse.<br />

Colonies on MEA reaching 48–49 mm diam after 14 d at 25 ºC,<br />

grey-olivaceous to olivaceous-grey, velvety, olivaceous-grey to<br />

138


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 32. <strong>Cladosporium</strong> ramotenellum (CPC 12043). A, C. Macronematous conidiophore. B. Conidial chain. D. Micronematous conidiophore. E. Ramoconidia <strong>and</strong> conidia. Scale<br />

bars = 10 µm.<br />

Fig. 33. <strong>Cladosporium</strong> ramotenellum (CPC 12043). A. Survey of colony development showing a large bulbous “foot cell” that gives rise to conidiophores, which can be branched.<br />

B. Details of conidiophores showing secondary ramoconidia <strong>and</strong> conidia. <strong>The</strong> inset shows scar formation on a conidiophore. C. Conidiophore <strong>and</strong> several conidia. D. Details<br />

of ornamentation on conidia. Note the wide, but relatively low ornamentation units. E. A micrograph illustrating the organisation within a conidiophore. Scale bars A–D = 5 µm,<br />

E = 10 µm.<br />

www.studiesinmycology.org<br />

139


Schubert et al.<br />

Fig. 34. <strong>Cladosporium</strong> sinuosum (CPC 11839). Conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

iron-grey reverse, margin entire edge to undulate, radially furrowed,<br />

colourless, glabrous to feathery, aerial mycelium sparse, diffuse,<br />

growth flat with slightly elevated colony centre, distinctly wrinkled,<br />

prominent exudates not formed, abundantly sporulating. Colonies<br />

on OA attaining 40 mm diam after 14 d at 25 ºC, grey-olivaceous,<br />

margin entire edge, colourless or white, glabrous, aerial mycelium<br />

absent or sparse, growth flat, without exudates, sporulation<br />

profuse.<br />

Specimens examined: Slovenia, Ljubljana, isolated from an air conditioning system<br />

(bathroom), 2004, M. Butala, <strong>CBS</strong> 121627 = CPC 12047 = EXF-967; Sečovlje,<br />

isolated from hypersaline water from reverse ponds, salterns, 2005, P. Zalar, <strong>CBS</strong>-H<br />

19862, holotype, isotype HAL 2026 F, culture ex-type <strong>CBS</strong> 121628 = CPC 12043<br />

= EXF-454.<br />

Substrate <strong>and</strong> distribution: Hypersaline water, air; Slovenia.<br />

Notes: <strong>Cladosporium</strong> ramotenellum, which appears to be a<br />

saprobe in air <strong>and</strong> hypersaline water, morphologically resembles<br />

C. cladosporioides <strong>and</strong> C. tenellum K. Schub., Zalar, Crous &<br />

U. Braun, but is quite distinct from C. cladosporioides by having<br />

somewhat narrower conidiophores <strong>and</strong> conidia, 2–4(–5) µm<br />

wide, <strong>and</strong> 0–3-septate, always minutely verruculose conidia.<br />

<strong>Cladosporium</strong> tenellum, a newly introduced species (see below)<br />

isolated from hypersaline water <strong>and</strong> plant material, possesses<br />

conidiophores with numerous conidiogenous loci, usually crowded<br />

towards the apex forming sympodial clusters of pronounced scars,<br />

<strong>and</strong> shorter <strong>and</strong> somewhat wider, 0–1(–2)-septate conidia, 3–20(–<br />

28) × (2.5–)3–5(–6) µm. Besides these morphological differences,<br />

C. ramotenellum is faster growing in culture than C. tenellum.<br />

<strong>Cladosporium</strong> arthrinioides Thüm. & Beltr. <strong>and</strong> C. hypophyllum<br />

Fuckel are also close to C. ramotenellum, but C. arthrinioides,<br />

known from Italy on leaves of Bougainvillea spectabilis, deviates in<br />

having shorter <strong>and</strong> wider, 0–1(–2)-septate, mostly smooth conidia<br />

(2–18 × 2–6.5 µm) which become larger <strong>and</strong> more frequently<br />

septate with age (up to 32 µm long <strong>and</strong> with up to four septa);<br />

<strong>and</strong> C. hypophyllum occurring in Europe on leaves of Ulmus<br />

minor differs in having often mildly to distinctly geniculate-sinuous,<br />

sometimes subnodulose conidiophores <strong>and</strong> shorter <strong>and</strong> somewhat<br />

wider, 0–1(–3)-septate conidia, 4–17(–19) × 2–5 µm, becoming<br />

distinctly swollen, darker, longer <strong>and</strong> wider with age, 5–7 µm, with<br />

the septa often being constricted (Schubert 2005).<br />

140


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 35. <strong>Cladosporium</strong> sinuosum (CPC 11839). A–D. Conidiophores. E–F. Conidia. Scale bars = 10 µm.<br />

<strong>Cladosporium</strong> sinuosum K. Schub., C.F. Hill, Crous & U. Braun,<br />

sp. nov. MycoBank MB504578. Figs 34–35.<br />

Etymology: Refers to the usually distinctly sinuous conidiophores.<br />

Differt a Cladosporio herbaro conidiophoris distincte sinuosis, conidiis solitariis vel<br />

breve catenatis, catenis non ramosis, echinulatis.<br />

Mycelium sparingly branched, 1–7 µm wide, septate, not<br />

constricted at the septa, subhyaline to pale brown, smooth to<br />

minutely verruculose, walls unthickened or slightly thickened,<br />

sometimes with small swellings. Conidiophores arising laterally<br />

from plagiotropous hyphae or terminally from ascending hyphae,<br />

erect, more or less straight to flexuous, often once or several times<br />

slightly to distinctly geniculate-sinuous, sometimes even zigzag-like,<br />

nodulose with small to large lateral shoulders, shoulders somewhat<br />

distant from each other or in close succession giving them a knotty/<br />

gnarled appearance, unbranched or once branched, 25–260 × 5–7<br />

µm, shoulders up to 10 µm wide, pluriseptate, septa sometimes<br />

in short succession, not constricted at the septa, pale brown to<br />

medium brown, smooth to minutely verruculose, walls thickened,<br />

often distinctly two-layered, up to 1 µm thick. Conidiogenous<br />

cells integrated, terminal or intercalary, often slightly to distinctly<br />

geniculate-sinuous, nodulose with small to large laterally swollen<br />

shoulders, 8–30 µm long, proliferation sympodial, with a single<br />

or up to three conidiogenous loci, usually confined to lateral<br />

www.studiesinmycology.org<br />

shoulders, protuberant, often denticle-like or on the top of short<br />

cylindrical stalk-like prolongations, 1.2–2(–2.2) µm diam, mainly<br />

2 µm, somewhat thickened <strong>and</strong> darkened-refractive, dome often<br />

slightly higher than the surrounding rim. Conidia solitary or in short<br />

unbranched chains with up to three conidia, straight, obovoid, oval,<br />

broadly ellipsoid to subcylindrical or sometimes clavate (broader at<br />

the apex), 9–21 × (5–)6–8 µm [av. ± SD, 14.5 (± 2.5) × 6.6 (± 0.7)<br />

µm], 0–1-septate, not constricted at the septa, septum more or less<br />

median, pale greyish brown, densely echinulate, spines up to 1 µm<br />

long, walls thickened, apex mostly broadly rounded or sometimes<br />

attenuated, towards the base mostly distinctly attenuated forming a<br />

peg-like prolongation, up to 2 µm long, hila protuberant, 1.2–2 µm<br />

diam, mainly 2 µm, somewhat thickened <strong>and</strong> darkened-refractive;<br />

microcyclic conidiogenesis not observed.<br />

Cultural characteristics: Colonies on PDA attaining 20 mm<br />

diam after 14 d at 25 ºC, pale olivaceous-grey due to abundant<br />

aerial mycelium, olivaceous-grey towards margins, iron-grey<br />

to olivaceous-black reverse, margin regular, entire edge, aerial<br />

mycelium abundant, cottony, dense, high, growth regular, low<br />

convex, radially furrowed in the centre, growing deep into the agar,<br />

with age numerous small to large prominent exudates, sporulation<br />

sparse. Colonies on MEA attaining 16 mm diam after 14 d at 25 ºC,<br />

white to pale smoke-grey, fawn reverse, velvety, margin undulate,<br />

glabrous, aerial mycelium abundant, dense, high, fluffy, growth<br />

raised with elevated colony centre, laterally furrowed, without<br />

141


Schubert et al.<br />

Fig. 36. <strong>Cladosporium</strong> spinulosum (CPC 12040). A. Overview on agar surface with conidiophores arising from the surface. <strong>The</strong> spore clusters on the conidiophore are very<br />

compact. Note several simple, tubular conidiophore ends. <strong>The</strong> inset shows details of a conidium showing two pronounced hila <strong>and</strong> a unique, very distinct ornamentation on the<br />

cell wall. B. Conidiophore with globose or subsphaerical secondary ramoconidia <strong>and</strong> conidia. Note the newly forming cells <strong>and</strong> hila. C. Two conidiophores. D. Details of spores<br />

<strong>and</strong> spore formation. E. <strong>The</strong> end of a conidiophore <strong>and</strong> two scars. Scale bars: A = 20 µm, A (inset) = 1 µm, B, D–E = 5 µm, C = 10 µm.<br />

prominent exudates. Colonies on OA attaining 18 mm diam after 14<br />

d at 25 ºC, olivaceous, white to pale olivaceous-grey in the centre<br />

due to abundant aerial mycelium, olivaceous-grey reverse, margin<br />

white, entire edge, glabrous, aerial mycelium loose to dense, high,<br />

fluffy to felty, growth flat to low convex, regular, without prominent<br />

exudates, sporulating.<br />

Specimen examined: New Zeal<strong>and</strong>, Te Anau, isolated from leaves of Fuchsia<br />

excorticata (Onagraceae), 31 Jan. 2005, A. Blouin, Hill 1134A, <strong>CBS</strong>-H 19863,<br />

holotype, culture ex-type <strong>CBS</strong> 121629 = CPC 11839 = ICMP 15819.<br />

Substrate <strong>and</strong> distribution: On living leaves of Fuchsia excorticata;<br />

New Zeal<strong>and</strong>.<br />

Notes: This new species is well characterised by its slightly to<br />

distinctly geniculate-sinuous, often zigzag-like conidiophores <strong>and</strong><br />

its conidia formed solitary or rarely in short unbranched chains <strong>and</strong><br />

is therefore morphologically not comparable with any of the species<br />

described until now. Most <strong>Cladosporium</strong> species with conidia usually<br />

formed solitary or in short unbranched chains have previously<br />

been treated as species of the <strong>genus</strong> Heterosporium Klotzsch ex<br />

Cooke, now considered to be synonymous with <strong>Cladosporium</strong>. All<br />

of them, including the newly introduced C. arthropodii K. Schub.<br />

& C.F. Hill from New Zeal<strong>and</strong>, which also belongs to this species<br />

complex (Braun et al. 2006), possess very large <strong>and</strong> wide, often<br />

pluriseptate conidia quite distinct from those of C. sinuosum (David<br />

1997). <strong>Cladosporium</strong> alopecuri (Ellis & Everh.) U. Braun, known<br />

from the U.S.A. on Alopecurus geniculatus is also quite different by<br />

having larger <strong>and</strong> wider conidia, 20–40 × 7–13(–15) µm, <strong>and</strong> wider<br />

conidiogenous loci <strong>and</strong> conidial hila, 3.5–5 µm diam (Braun 2000).<br />

<strong>Cladosporium</strong> herbarum is superficially <strong>similar</strong> but the<br />

conidiophores of the latter species are sometimes only slightly<br />

geniculate-sinuous but never zigzag-like <strong>and</strong> the verruculose to<br />

verrucose conidia are frequently formed in unbranched or branched<br />

chains.<br />

<strong>Cladosporium</strong> spinulosum Zalar, de Hoog & Gunde-Cimerman,<br />

Studies in Mycology 58: 180. 2007 – this volume. Fig. 36.<br />

Note: This new species is described <strong>and</strong> illustrated in Zalar et al.<br />

(2007 – this volume).<br />

142


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 37. <strong>Cladosporium</strong> subinflatum (CPC 12041). Macro- <strong>and</strong> micronematous conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

<strong>Cladosporium</strong> subinflatum K. Schub., Zalar, Crous & U. Braun,<br />

sp. nov. MycoBank MB504579. Figs 37–39.<br />

Etymology: Refers to its nodulose conidiophores.<br />

Differt a Cladosporio bruhnei conidiophoris cum nodulis angustioribus, 3–6.5 µm<br />

latis, conidiis brevioribus, 4–17(–22) µm longis, spinulosis, cum spinulis ad 0.8 µm<br />

longis; et a Cladosporio spinuloso conidiophoris nodulosis, conidiis spinulosis, cum<br />

spinulis brevioribus, ad 0.8 longis, locis conidiogenis et hilis latioribus, (0.5–)1–2<br />

µm latis.<br />

Mycelium unbranched or occasionally branched, 1.5–3 µm wide,<br />

later more frequently branched <strong>and</strong> wider, up to 7 µm wide,<br />

septate, not constricted at the septa, hyaline or subhyaline, almost<br />

smooth to somewhat verruculose or irregularly rough-walled, walls<br />

unthickened. Conidiophores mainly macronematous, sometimes<br />

also micronematous, arising terminally from ascending hyphae<br />

or laterally from plagiotropous hyphae, erect or subdecumbent,<br />

straight or flexuous, sometimes bent, cylindrical, nodulose, usually<br />

with small head-like swellings, sometimes swellings also on a<br />

lower level or intercalary, occasionally geniculate, unbranched,<br />

occasionally branched, (5–)10–270 × (1.5–)2.5–4.5(–5.5) µm,<br />

www.studiesinmycology.org<br />

swellings 3–6.5 µm wide, aseptate or with few septa, not constricted<br />

at the septa, pale brown, pale olivaceous-brown or somewhat<br />

reddish brown, smooth, usually verruculose or irregularly roughwalled<br />

<strong>and</strong> paler, subhyaline towards the base, walls thickened,<br />

sometimes appearing even two-layered, up to 1 µm thick.<br />

Conidiogenous cells integrated, usually terminal or conidiophores<br />

reduced to conidiogenous cells, cylindrical, nodulose, usually with<br />

small head-like swellings with loci confined to swellings, sometimes<br />

geniculate, 5–42 µm long, proliferation sympodial, with several loci,<br />

up to four situated at nodules or on lateral swellings, protuberant,<br />

conspicuous, denticulate, (0.8–)1–2 µm diam, thickened <strong>and</strong><br />

darkened-refractive. Conidia catenate, in branched chains, more<br />

or less straight, numerous globose <strong>and</strong> subglobose conidia, ovoid,<br />

obovoid, broadly ellipsoid to cylindrical, 4–17(–22) × (2.5–)3.5–<br />

5.5(–7) µm [av. ± SD, 11.7 (± 4.6) × 4.5 (± 0.8) µm], 0–1(–2)-<br />

septate, not constricted at septa, pale brown or pale olivaceousbrown,<br />

ornamentation variable, mainly densely verruculose to<br />

echinulate (loosely muricate under SEM), spines up to 0.8 µm<br />

high, sometimes irregularly verrucose with few scattered tubercles<br />

143


Schubert et al.<br />

Fig. 38. <strong>Cladosporium</strong> subinflatum (CPC 12041). A–C. Macronematous conidiophores. D–E. Conidia. Scale bar = 10 µm.<br />

Fig. 39. <strong>Cladosporium</strong> subinflatum<br />

(CPC 12041). A–G. Images of an 11-<br />

d-old culture on SNA. A. Overview of<br />

colony with clusters of conidia <strong>and</strong><br />

aerial hyphae. Many of the hyphae<br />

have a collapsed appearance. B.<br />

Detail of colony with conidiophores,<br />

conidia <strong>and</strong> aerial hyphae that<br />

are partly collapsed. C. Detail of a<br />

conidiophore end <strong>and</strong> a secondary<br />

ramoconidium. Note the scars at the<br />

end of the conidiophore. D. Details<br />

of conidia <strong>and</strong> ornamentation.<br />

<strong>The</strong> ornamentation consists out of<br />

markedly defined units, which have<br />

a relatively large distance from each<br />

other. Note the hilum on the right<br />

conidium. E. Conidiophore with<br />

large scars <strong>and</strong> conidia. F. Different<br />

blastoconidia with very early stages<br />

of new spore formation in the middle<br />

of the picture. G. Pattern of spore<br />

development. Scale bars: A = 20<br />

µm, B, E–G = 5 µm, C = 10 µm, D<br />

= 2 µm.<br />

144


<strong>Cladosporium</strong> herbarum species complex<br />

or irregularly echinulate, walls unthickened or slightly thickened,<br />

apex rounded or slightly attenuated towards apex <strong>and</strong> base, hila<br />

conspicuous, protuberant, denticulate, 0.5–2 µm diam, thickened<br />

<strong>and</strong> darkened-refractive; microcyclic conidiogenesis observed.<br />

Cultural characteristics: Colonies on PDA attaining 29 mm diam<br />

after 14 d at 25 ºC, olivaceous-black to olivaceous-grey towards<br />

margin, margin regular, entire edge, narrow, colourless to white,<br />

glabrous to feathery, aerial mycelium formed, fluffy, mainly near<br />

margins, growth flat, somewhat folded in the colony centre, deep<br />

into the agar, few prominent exudates formed with age, sporulation<br />

profuse. Colonies on MEA attaining 25 mm diam after 14 d at 25<br />

ºC, olivaceous-grey to olivaceous due to abundant sporulation<br />

in the colony centre, pale greenish grey towards margin, irongrey<br />

reverse, velvety to powdery, margin crenate, narrow, white,<br />

glabrous, radially furrowed, aerial mycelium diffuse, growth convex<br />

with papillate surface, wrinkled colony centre, without prominent<br />

exudates, sporulation profuse. Colonies on OA attaining 26 mm<br />

diam after 14 d at 25 ºC, olivaceous, iron-grey to greenish black<br />

reverse, growth flat, deep into the agar, with a single exudate,<br />

abundantly sporulating.<br />

Specimen examined: Slovenia, Sečovlje, isolated from hypersaline water from<br />

crystallization ponds, salterns, 2005, S. Sonjak, <strong>CBS</strong>-H 19864, holotype, isotype<br />

HAL 2027 F, culture ex-type <strong>CBS</strong> 121630 = CPC 12041 = EXF-343.<br />

Substrate <strong>and</strong> distribution: Hypersaline water; Slovenia.<br />

Notes: <strong>Cladosporium</strong> subinflatum, an additional saprobic species<br />

isolated from hypersaline water, was at first identified as C.<br />

spinulosum, but proved to be both morphologically as well as<br />

phylogenetically distinct from the latter species in having somewhat<br />

wider [(1.5–)2.5–4.5(–5.5) µm], nodulose macronematous<br />

conidiophores with conidiogenous loci confined to swellings, wider<br />

conidiogenous loci <strong>and</strong> hila, (0.8–)1–2 µm, <strong>and</strong> spiny conidia<br />

with shorter spines than in C. spinulosum (up to 0.8 µm versus<br />

0.5–1.3 µm long) (Zalar et al. 2007). With its narrow, nodulose<br />

macronematous conidiophores <strong>and</strong> catenate conidia, C. bruhnei<br />

is morphologically also <strong>similar</strong> but differs by having conidiophores<br />

with wider swellings, (4–)5–8 µm, <strong>and</strong> longer conidia 4–24(–31)<br />

µm, rarely up to 40 µm long which are minutely verruculose to<br />

verrucose but not spiny.<br />

Fig. 40. <strong>Cladosporium</strong> subtilissimum (<strong>CBS</strong> 113754). Macro- <strong>and</strong> micronematous conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

www.studiesinmycology.org<br />

145


Schubert et al.<br />

Fig. 41. <strong>Cladosporium</strong> subtilissimum (<strong>CBS</strong> 113754). A–C. Macronematous conidiophores. D. Conidial chain. E. Micronematous conidiophore. F–G. Conidia. Scale bars = 10<br />

µm.<br />

<strong>Cladosporium</strong> subtilissimum K. Schub., Dugan, Crous & U.<br />

Braun, sp. nov. MycoBank MB504580. Figs 40–42.<br />

Etymology: Refers to its narrow conidiophores <strong>and</strong> conidia.<br />

Differt a Cladosporio cladosporioide conidiophoris et conidiis semper asperulatis ad<br />

verruculosis, conidiis 0–1(–2)-septatis.<br />

Mycelium unbranched or sparingly branched, 1–5 µm wide, septate,<br />

without swellings <strong>and</strong> constrictions, hyaline to subhyaline or pale<br />

brown, smooth to minutely verruculose, walls unthickened or almost<br />

so, protoplasm somewhat guttulate or granular. Conidiophores<br />

macronematous <strong>and</strong> micronematous, arising laterally from<br />

plagiotropous hyphae or terminally from ascending hyphae, erect,<br />

straight to slightly flexuous, filiform to cylindrical-oblong, nonnodulose,<br />

sometimes geniculate towards the apex, unbranched or<br />

once branched, branches short to somewhat longer, usually formed<br />

below a septum, sometimes only short, denticle-like or conical,<br />

25–140 × 2–4 µm, 0–4-septate, not constricted at the septa,<br />

subhyaline to pale brown, almost smooth, minutely verruculose<br />

to verruculose, sometimes irregularly rough-walled in the lower<br />

part, walls unthickened or slightly thickened, protoplasm guttulate<br />

or somewhat granular. Conidiogenous cells integrated, terminal<br />

or pleurogenous, sometimes also intercalary, filiform to narrowly<br />

cylindrical, non-nodulose, sometimes geniculate, 14–57 µm long,<br />

with usually sympodial clusters of pronounced conidiogenous loci<br />

at the apex or on a lower level, denticle-like or situated on short<br />

lateral prolongations, up to five loci, intercalary conidiogenous<br />

cells usually with a short denticle-like lateral outgrowth below a<br />

septum, protuberant, denticulate, somewhat truncate, 1.2–2 µm<br />

diam, thickened <strong>and</strong> darkened-refractive. Ramoconidia sometimes<br />

occurring, conidiogenous cells seceding at one of the upper septa<br />

of the conidiophore <strong>and</strong> behaving like conidia, filiform or cylindrical,<br />

20–40(–55) µm long, 1.5–4 µm wide, 0–1-septate, concolorous<br />

with conidiophores, not attenuated towards apex <strong>and</strong> base, base<br />

broadly truncate, non-cladosporioid, without any dome <strong>and</strong> raised<br />

rim, 2–3.5 µm wide, neither thickened nor darkened, sometimes<br />

slightly refractive. Conidia catenate, in branched chains, up to 12<br />

or even more in a chain, straight, small terminal conidia numerous,<br />

subglobose, narrowly obovoid, limoniform or fusiform, 4–9 × 2–3.5<br />

µm [av. ± SD, 6.4 (± 1.5) × 2.8 (± 0.4) µm], with up to three distal<br />

scars, aseptate, hila (0.5–)0.8–1 µm diam, intercalary conidia<br />

narrowly ellipsoid, fusiform to subcylindrical, 9–18 × 3–4(–6) µm<br />

[av. ± SD, 13.0 (± 2.5) × 3.8 (± 0.3) µm], 0(–1)-septate, hila 1–1.2(–<br />

1.8) µm diam, with up to four distal scars, secondary ramoconidia<br />

ellipsoid, fusiform or subcylindrical, (13–)17–32(–37) × 3–5(–6) µm<br />

[av. ± SD, 21.4 (± 4.4) × 4.1 (± 0.5) µm], 0–1(–2)-septate, septum<br />

median or somewhat in the lower half, usually not constricted at<br />

the septa, with up to six distal hila crowded at the apex, hila (1.2–)<br />

1.5–2(–2.5) µm diam, apex often somewhat laterally enlarged<br />

or prolonged with hila crowded there, very pale or pale brown or<br />

olivaceous-brown, minutely verruculose to verruculose (granulate<br />

under SEM), walls unthickened or only slightly thickened, often<br />

slightly attenuated towards apex <strong>and</strong> base, protoplasm often<br />

guttulate or granular, hila protuberant, denticulate, (0.5–)0.8–2(–<br />

2.2) µm diam, thickened <strong>and</strong> darkened-refractive; microcyclic<br />

conidiogenesis occasionally observed.<br />

Cultural characteristics: Colonies on PDA attaining 24 mm diam<br />

after 14 d at 25 °C, grey-olivaceous to olivaceous, olivaceousgrey,<br />

iron-grey or olivaceous-black reverse, velvety, margin<br />

regular, entire edge, white or pale greenish olivaceous, glabrous<br />

to feathery, aerial mycelium sparse, only few areas with abundant<br />

146


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 42. <strong>Cladosporium</strong> subtilissimum (<strong>CBS</strong> 113754). A. Overview on the organisation of spore formation. <strong>The</strong> micrograph shows a large basal secondary ramoconidium which<br />

has chains of secondary ramoconidia, intercalary <strong>and</strong> small terminal conidia. <strong>The</strong> conidia are formed in rows of often three cells. Note the size difference in the different cells.<br />

B. Conidiophore showing very pronounced scars that almost appear as branches. C. Detail of (A), illustrating the scar formation between the cells. D. Conidia during different<br />

stages of formation. E. Details of pronounced hila, <strong>and</strong> prominent ornamentation on secondary ramoconidia with the central dome-formed area. F. Different conidia <strong>and</strong> hila.<br />

Scale bars: A = 10 µm, B–D, F = 5 µm, E = 2 µm.<br />

mycelium, diffuse, growth regular, flat or with a raised <strong>and</strong> wrinkled<br />

colony centre, radially furrowed, effuse, usually without prominent<br />

exudates, with age several exudates formed, sporulation profuse,<br />

colonies consisting of two kinds of conidiophores, short <strong>and</strong> a few<br />

longer ones. Colonies on MEA reaching 25 mm diam after 14 d<br />

at 25 °C, greenish olivaceous to grey-olivaceous in the centre,<br />

olivaceous-grey to iron-grey reverse, velvety, margin entire edge,<br />

crenate or umbonate, narrow, pale greenish olivaceous, sometimes<br />

radially furrowed, aerial mycelium absent or sparse, growth low<br />

convex with distinctly wrinkled colony centre, without prominent<br />

exudates, abundantly sporulating. Colonies on OA attaining 25 mm<br />

diam after 14 d at 25 °C, dark grey-olivaceous to olivaceous due to<br />

profuse sporulation, iron-grey reverse, sometimes releasing some<br />

olivaceous-buff pigments into the agar, velvety, margin regular,<br />

entire edge or crenate, narrow, colourless or white, glabrous or<br />

feathery, aerial mycelium sparse, growth flat with slightly raised<br />

colony centre, prominent exudates lacking, sporulation profuse.<br />

www.studiesinmycology.org<br />

Specimens examined: Slovenia, Sečovlje, isolated from hypersaline water from<br />

salterns (reserve pond), 2005, P. Zalar, CPC 12044 = EXF-462. U.S.A., isolated<br />

from bing cherry fruits, F. Dugan, <strong>CBS</strong> 113753; isolated from a grape berry, F.<br />

Dugan, wf 99-2-9 sci 1, <strong>CBS</strong>-H 19865, holotype, isotype HAL 2028 F, culture extype<br />

<strong>CBS</strong> 113754.<br />

Excluded strains within the subtilissimum complex: Argentina,<br />

isolated from Pinus ponderosa (Pinaceae), 2005, A. Greslebin,<br />

CPC 12484, CPC 12485. U.S.A., isolated from grape berry, F.<br />

Dugan, <strong>CBS</strong> 113741, <strong>CBS</strong> 113742; isolated from grape bud, F.<br />

Dugan, <strong>CBS</strong> 113744.<br />

Substrate <strong>and</strong> distribution: Plant material <strong>and</strong> hypersaline water;<br />

Slovenia, U.S.A.<br />

Notes: <strong>Cladosporium</strong> cladosporioides is morphologically comparable<br />

with the new species but deviates in having usually smooth<br />

conidiophores <strong>and</strong> conidia, with the conidia being mainly aseptate.<br />

C. subtilissimum is represented by three isolates of different origins<br />

147


Schubert et al.<br />

Fig. 43. <strong>Cladosporium</strong> tenellum (CPC 12053). Macro- <strong>and</strong> micronematous conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

Fig. 44. <strong>Cladosporium</strong> tenellum (CPC 12053). A–C, E. Macronematous conidiophore. D. Micronematous conidiophore. F. Ramoconidium <strong>and</strong> conidia. Scale bars = 10 µm.<br />

148


<strong>Cladosporium</strong> herbarum species complex<br />

Differt a Cladosporio cladosporioide conidiophoris et conidiis semper asperulatis,<br />

locis conidiogenis apicalibus, numerosis, hilis quoque numerosis, conidiophoris<br />

angustioribus, (1–)1.5–3.5(–4) µm latis; et a Cladosporio subtilissimo loci<br />

conidiogenis et hilis apicalibus, numerosis, angustioribus, saepe 1–1.5 µm latis,<br />

conidiis minutis numerosis, saepe globosis.<br />

Fig. 45. <strong>Cladosporium</strong> tenellum (CPC 12053). A. A bird’s eye view of a colony of<br />

C. tenellum with its very characteristic bundles of aerial hyphae. Numerous conidia<br />

are visible, formed on simple conidiophores. B. Hyphae that run on the agar surface<br />

give rise to conidiophores <strong>and</strong> numerous conidia, that are relatively rounded. C.<br />

Conidiophore ends are rather simple <strong>and</strong> have large scars. D. Hila on a secondary<br />

ramoconidium with non-ornamented area. E. Detail of the prominent ornamentation<br />

on a secondary ramoconidium. Scale bars: A = 20 µm, B = 10 µm, C, E = 2 µm,<br />

D = 5 µm.<br />

<strong>and</strong> substrates. Besides these strains, several additional isolates<br />

listed under excluded strains are morphologically indistinguishable<br />

from C. subtilissimum in culture, but genetically different, clustering<br />

in various subclades. <strong>The</strong>y are indicated as <strong>Cladosporium</strong> sp. in<br />

the tree (Fig. 3).<br />

<strong>Cladosporium</strong> tenellum K. Schub., Zalar, Crous & U. Braun, sp.<br />

nov. MycoBank MB504581. Figs 43–45.<br />

Etymology: Refers to its narrow conidiophores <strong>and</strong> conidia.<br />

www.studiesinmycology.org<br />

Mycelium sparingly branched, 1–3 µm wide, septate, septa often<br />

not very conspicuous, not constricted at the septa, sometimes<br />

slightly swollen, subhyaline, smooth, walls unthickened.<br />

Conidiophores macronematous <strong>and</strong> micronematous, solitary,<br />

arising terminally or laterally from plagiotropous or ascending<br />

hyphae, erect or subdecumbent, almost straight to more or less<br />

flexuous, cylindrical, sometimes geniculate towards the apex, but<br />

not nodulose, sometimes with short lateral prolongations at the<br />

apex, unbranched to once or twice branched (angle usually 30–45°<br />

degree, sometimes up to 90°), branches usually below a septum,<br />

6–200 × (1–)2–4(–5) µm, septate, septa not very conspicuous,<br />

not constricted at the septa, subhyaline to pale brown, almost<br />

smooth to usually asperulate, walls unthickened or almost so.<br />

Conidiogenous cells integrated, terminal or intercalary, sometimes<br />

conidiophores reduced to conidiogenous cells, cylindrical,<br />

sometimes geniculate, non-nodulose, 6–40 µm long, proliferation<br />

sympodial, with several conidiogenous loci often crowded at the<br />

apex <strong>and</strong> sometimes also at a lower level, situated on small lateral<br />

shoulders, unilateral swellings or prolongations, with up to 6(–10)<br />

denticulate loci, forming sympodial clusters of pronounced scars,<br />

intercalar conidiogenous cells with short or somewhat long lateral<br />

outgrowths, short denticle-like or long branches with several scars<br />

at the apex, usually below a septum, loci protuberant, 1–1.5(–2) µm<br />

diam, thickened <strong>and</strong> darkened-refractive. Ramoconidia sometimes<br />

occurring, cylindrical, up to 32 µm long, 2.5–4 µm wide, with a<br />

broadly truncate, unthickened base, about 2 µm wide. Conidia<br />

catenate, formed in branched chains, straight, small terminal<br />

conidia globose, subglobose, ovoid, oval, 3–6 × 2.5–3.5 µm [av.<br />

± SD, 4.5 (± 1.3) × 2.8 (± 0.4) µm], aseptate, asperulate, with 0–2<br />

distal hila, intercalary conidia <strong>and</strong> secondary ramoconidia ellipsoidovoid,<br />

ellipsoid to subcylindrical, 3.5–20(–28) × (2.5–)3–5(–6) µm<br />

[av. ± SD, 12.4 (± 5.4) × 4.1 (± 0.7) µm], 0–1-septate, rarely with<br />

up to three septa, sometimes slightly constricted at the septa,<br />

subhyaline, pale brown to medium olivaceous-brown, asperulate<br />

or verruculose (muricate, granulate or colliculate under SEM),<br />

walls unthickened or slightly thickened, apex rounded or slightly<br />

to distinctly attenuated towards apex <strong>and</strong> base, often forming<br />

several apical hila, up to 7(–9), crowded, situated on small lateral<br />

outgrowths giving them a somewhat irregular appearance, hila<br />

protuberant, 0.5–1.5 µm diam, thickened <strong>and</strong> darkened-refractive;<br />

microcyclic conidiogenesis sometimes occurring.<br />

Cultural characteristics: Colonies on PDA reaching 27–34 mm diam<br />

after 14 d at 25 ºC, smoke-grey, grey-olivaceous to olivaceous-grey,<br />

olivaceous-grey to iron-grey reverse, velvety to powdery, margin<br />

regular, entire edge, narrow, colourless to white, aerial mycelium<br />

absent or sparingly formed, felty, whitish, growth regular, flat,<br />

radially furrowed, with folded <strong>and</strong> elevated colony centre, deep into<br />

the agar, with age forming few to numerous prominent exudates,<br />

sporulation profuse, few high conidiophores formed. Colonies on<br />

MEA reaching 25–44 mm diam after 14 d at 25 ºC, olivaceous-grey<br />

to olivaceous- or iron-grey due to abundant sporulation in the colony<br />

centre, velvety, margin regular, entire edge, narrow, colourless,<br />

white to pale olivaceous-grey, aerial mycelium loose, diffuse,<br />

growth convex with papillate surface, radially furrowed, wrinkled,<br />

without prominent exudates, sporulating. Colonies on OA reaching<br />

23–32 mm diam after 14 d at 25 ºC, grey-olivaceous, olivaceous-<br />

149


Schubert et al.<br />

Fig. 46. <strong>Cladosporium</strong> variabile (CPC 12751). Macro- <strong>and</strong> micronematous conidiophores <strong>and</strong> conidia. Scale bar = 10 µm. K. Schubert del.<br />

grey to olivaceous due to abundant sporulation in the colony centre,<br />

olivaceous- or iron-grey reverse, velvety, margin regular, entire<br />

edge, narrow, colourless or white, aerial mycelium sparse, diffuse,<br />

floccose, growth flat to low convex, radially furrowed, wrinkled,<br />

without prominent exudates, sporulation profuse.<br />

Specimens examined: Israel, Eilat, isolated from hypersaline water from salterns,<br />

2004, N. Gunde-Cimerman, <strong>CBS</strong> 121633 = CPC 12051 = EXF-1083; Ein Bokek,<br />

isolated from hypersaline water of the Dead Sea, 2004, M. Ota, <strong>CBS</strong>-H 19866,<br />

holotype, isotype HAL 2029 F, culture ex-type <strong>CBS</strong> 121634 = CPC 12053 = EXF-<br />

1735. U.S.A., Seattle, University of Washington campus, isolated from Phyllactinia<br />

sp. (Erysiphaceae) on leaves of Corylus sp. (Corylaceae), 16 Sep. 2004, D. Glawe,<br />

CPC 11813.<br />

Substrates <strong>and</strong> distribution: Hypersaline water <strong>and</strong> plant material;<br />

Israel, U.S.A.<br />

Notes: <strong>Cladosporium</strong> subtilissimum <strong>and</strong> C. cladosporioides are<br />

morphologically comparable with the new species C. tenellum, but<br />

C. cladosporioides deviates in having usually smooth conidiophores<br />

<strong>and</strong> conidia with only few conidiogenous loci <strong>and</strong> conidial hila crowded<br />

at the apex <strong>and</strong> somewhat wider conidiophores, 3–5(–6) µm; <strong>and</strong><br />

in C. subtilissimum the small terminal conidia are not globose but<br />

rather narrowly obovoid to limoniform, the conidiogenous loci <strong>and</strong><br />

conidial hila are somewhat wider, (0.5–)0.8–2(–2.2) µm, <strong>and</strong> at the<br />

apices of conidiophores <strong>and</strong> conidia only few scars are formed.<br />

<strong>Cladosporium</strong> ramotenellum, which morphologically also<br />

resembles C. tenellum, possesses longer <strong>and</strong> narrower, 0–3-septate<br />

conidia, 2.5–35 × 2–4(–5) µm, but forms only few conidiogenous<br />

loci <strong>and</strong> conidial hila at the apices of conidiophores <strong>and</strong> conidia.<br />

<strong>Cladosporium</strong> variabile (Cooke) G.A. de Vries, Contr. Knowl.<br />

Genus <strong>Cladosporium</strong>: 85. 1952. Figs 46–48.<br />

Basionym: Heterosporium variabile Cooke, Grevillea 5(35): 123.<br />

1877.<br />

≡ Helminthosporium variabile Cooke, Fungi Brit. Exs. Ser. 1, No. 360.<br />

1870, nom. inval.<br />

= <strong>Cladosporium</strong> subnodosum Cooke, Grevillea 17(83): 67. 1889.<br />

150


<strong>Cladosporium</strong> herbarum species complex<br />

Fig. 47. <strong>Cladosporium</strong> variabile <strong>and</strong> its teleomorph Davidiella variabile (CPC 12751). A–C. Macronematous conidiophores. D, F. Micronematous conidiophores. E, G–H.<br />

Conidia. I. Twisted aerial mycelium. J. Ascomata formed on nettle stem in culture. K. Surface view of ascomal wall of textura epidermoidea. L–M. Asci. N–P. Ascospores. Q.<br />

Ascus with a sheath. Scale bars A, D, G–J, K–N = 10 µm, J = 250 µm.<br />

www.studiesinmycology.org<br />

151


Schubert et al.<br />

Fig. 48. <strong>Cladosporium</strong> variabile (CPC 12753). A. Survey of hyphae that grow on the agar surface. Some of the fungal cells have a swollen appearance <strong>and</strong> could develop into<br />

a “foot cell” that gives rise to a conidiophore. B. A number of aerial hyphae obstruct the swollen, large structures on the agar surface, which give rise to conidiophores. Some<br />

of them appear ornamented. C. A series of conidia formed on a conidiophore (bottom of the micrograph). D. Detail of the ornamented conidia. <strong>The</strong> ornamentations are isolated<br />

<strong>and</strong> dispersed. Note also the ornamentation-free scar zone <strong>and</strong> the hilum of the left cell. E. Two conidia behind an aerial hypha. F. Two conidiophores forming secondary<br />

ramoconidia. Note the bulbous shape of the spore-forming apparatus. This micrograph is from an uncoated sample. Scale bars: A–C, F = 10 µm, D = 2 µm, E = 5 µm.<br />

Teleomorph: Davidiella variabile Crous, K. Schub. & U. Braun, sp.<br />

nov. MycoBank MB504583.<br />

Davidiellae tassianae similis, sed ascosporis maioribus, (22–)26–30(–35) × (7–)7.5–<br />

8(–9) µm, et ascis latioribus, plus quam 18 µm.<br />

Ascomata pseudothecial, black, superficial, situated on a small<br />

stroma, globose, up to 250 µm diam, with 1–3 ostiolate necks;<br />

ostioles periphysate, with apical periphysoids present; wall<br />

consisting of 3–6 layers of dark brown textura angularis, textura<br />

epidermoidea in surface view. Asci fasciculate, bitunicate,<br />

subsessile, obovoid to broadly ellipsoid, straight to slightly curved,<br />

8-spored, 70–95 × 18–28 µm; with pseudoparenchymatal cells<br />

of the hamathecium persistent. Ascospores tri- to multiseriate,<br />

overlapping, hyaline, with irregular lumina, thick-walled, straight<br />

to slightly curved, fusoid-ellipsoidal with obtuse ends, widest near<br />

the middle of the apical cell, medianly 1-septate, not to slightly<br />

constricted at the septum, at times developing a second septum<br />

in each cell, several ascospores with persistent, irregular mucoid<br />

sheath, (22–)26–30(–35) × (7–)7.5–8(–9) µm.<br />

Mycelium immersed <strong>and</strong> superficial, irregularly branched, aerial<br />

mycelium twisted <strong>and</strong> spirally coiled, 1–3 µm wide, septate,<br />

sometimes with swellings or small lateral outgrowths, hyaline<br />

to subhyaline, smooth, walls unthickened, hyphae which give<br />

rise to conidiophores somewhat wider, 3–4.5 µm, subhyaline to<br />

pale brown, almost smooth to minutely verruculose, sometimes<br />

enveloped by a polysaccharide-like cover. Conidiophores usually<br />

macronematous, but also micronematous, arising terminally<br />

from ascending hyphae or laterally from plagiotropous hyphae.<br />

Macronematous conidiophores erect, more or less straight to<br />

flexuous, often distinctly geniculate-sinuous forming lateral<br />

shoulders or unilateral swellings, sometimes zigzag-like or<br />

somewhat coralloid, nodulose, swellings at first terminal, then<br />

becoming lateral due to sympodial proliferation, often as distinct<br />

lateral shoulders, unbranched, sometimes once branched, 6–<br />

180 × (2.5–)3–6 µm, swellings (3–)6–11 µm wide, septate, not<br />

constricted at the septa, pale to medium olivaceous-brown or<br />

brown, usually verruculose, walls somewhat thickened, about 1 µm<br />

thick, sometimes appearing to be two-layered. Conidiogenous cells<br />

integrated, terminal <strong>and</strong> intercalary, cylindrical, nodulose to nodose,<br />

152


<strong>Cladosporium</strong> herbarum species complex<br />

with a single or two swellings per cell, swellings apart from each<br />

other or formed in short succession, loci confined to swellings, up to<br />

six per node, protuberant, 1–2 µm diam, thickened <strong>and</strong> darkenedrefractive.<br />

Micronematous conidiophores erect, straight to slightly<br />

flexuous, unbranched, usually without swellings, filiform to narrowly<br />

cylindrical, sometimes only as short lateral outgrowths of hyphae,<br />

often almost indistinguishable from hyphae, up to 50 µm long,<br />

1.5–2.5(–3) µm wide, longer ones pluriseptate, septa appear to be<br />

somewhat more darkened, with very short cells, 4–12 µm long,<br />

subhyaline to pale brown, smooth, walls unthickened or almost so.<br />

Conidiogenous cells integrated, usually terminal, rarely intercalary,<br />

cylindrical, non-nodulose, with a single, two or few conidiogenous<br />

loci at the distal end, protuberant, up to 2 µm diam, thickened<br />

<strong>and</strong> darkened-refractive. Conidia catenate, in branched chains,<br />

straight, subglobose, obovoid, oval, broadly ellipsoid to cylindrical,<br />

sometimes clavate, 4–26(–30) × (3.5–)5–9(–10) µm [av. ± SD,<br />

16.8 (± 6.9) × 6.5 (± 1.4) µm], 0–3-septate, usually not constricted<br />

at the septa, septa becoming sinuous with age, often appearing to<br />

be darkened, pale to medium or even dark brown or olivaceousbrown,<br />

verruculose to densely verrucose or echinulate (granulate<br />

under SEM), walls slightly to distinctly thickened in larger conidia,<br />

apex <strong>and</strong> base broadly rounded, sometimes broadly truncate or<br />

somewhat attenuated, apex <strong>and</strong> base often appear to be darkened<br />

or at least refractive, hila protuberant to somewhat sessile (within<br />

the outer wall ornamentation), (0.8–)1–2 µm diam, thickened <strong>and</strong><br />

darkened-refractive; microcyclic conidiogenesis occurring.<br />

Cultural characteristics: Colonies on PDA attaining 29 mm diam<br />

after 14 d at 25 ºC, olivaceous to olivaceous-grey or iron-grey, irongrey<br />

or olivaceous-grey reverse, velvety to powdery, margin regular,<br />

entire edge to fimbriate, almost colourless, aerial mycelium whitish<br />

turning olivaceous-grey, sometimes reddish, greyish rose, woollyfelty,<br />

growth flat with elevated colony centre, somewhat folded or<br />

radially furrowed, with age forming several very small but prominent<br />

exudates, sporulation profuse. Colonies on MEA attaining 27 mm<br />

diam after 14 d at 25 ºC, olivaceous-grey to iron-grey, white to pale<br />

olivaceous-grey in the centre due to abundant aerial mycelium,<br />

velvety, margin very narrow, colourless, more or less entire edge,<br />

radially furrowed, aerial mycelium fluffy to floccose, dense, growth<br />

low convex with wrinkled <strong>and</strong> folded centre, without exudates,<br />

sporulation profuse. Colonies on OA attaining 25 mm diam after<br />

14 d at 25 ºC, iron-grey or olivaceous, margin regular, entire edge,<br />

narrow, white, glabrous, aerial mycelium whitish, at first mainly in<br />

the colony centre, high, dense, floccose, growth flat, abundantly<br />

sporulating, no exudates.<br />

Specimens examined: Great Britain, Wales, Montgomeryshire, Welshpool, Forden<br />

Vicarage, on Spinacia oleracea (Chenopodiaceae), J.E. Vize, Cooke, Fungi Brit. Exs.<br />

Ser. I, No. 360, K, holotype. U.S.A., Washington, isolated from Spinacia oleracea, 1<br />

Jan. 2003, L. DuToit, <strong>CBS</strong>-H 19867, epitype designated here of C. variabile <strong>and</strong> D.<br />

variabile, cultures ex-epitype <strong>CBS</strong> 121635 = CPC 12753, CPC 12751.<br />

Substrate <strong>and</strong> distribution: Leaf-spotting fungus on Spinacia<br />

oleracea; Asia (China, India, Iraq, Pakistan), Europe (Austria,<br />

Belgium, Cyprus, Denmark, France, Germany, Great Britain,<br />

Hungary, Italy, Montenegro, Netherl<strong>and</strong>s, Norway, Romania, Spain,<br />

Turkey), North America (U.S.A.).<br />

Literature: de Vries (1952: 85–88), Ellis (1971: 315), Ellis & Ellis<br />

(1985: 429), David (1995b; 1997: 94, 96–98), Ho et al. (1999:<br />

144).<br />

Notes: In vivo the conidia are usually longer, somewhat wider <strong>and</strong><br />

more frequently septate, (6.5–)10–45(–55) × (4.5–)6–14(–17) µm,<br />

www.studiesinmycology.org<br />

0–4(–5)-septate (Schubert 2005). In culture the dimensions tend to<br />

be smaller, which was already mentioned by de Vries (1952).<br />

This leaf-spotting fungus superficially resembles C.<br />

macrocarpum, but besides its pathogenicity to Spinacia, C.<br />

variabile differs from the latter species in having distinctly larger<br />

<strong>and</strong> more frequently septate conidia on the natural host, forming<br />

twisted <strong>and</strong> spirally coiled aerial mycelium in culture <strong>and</strong> in having<br />

lower growth rates in culture (29 mm after 14 d on PDA versus<br />

38 mm on average in C. macrocarpum). Furthermore, the conidial<br />

septa of C. variabile are often distinctly darkened, become sinuous<br />

with age <strong>and</strong> the apex <strong>and</strong> base of the conidia often appear to<br />

be distinctly darkened. A Davidiella teleomorph has not previously<br />

been reported for this species.<br />

<strong>The</strong> cladosporioides complex<br />

This species complex will be treated in an additional paper in<br />

this series, dealing with the epitypification of this common <strong>and</strong><br />

widespread species, <strong>and</strong> with numerous isolates identified <strong>and</strong><br />

deposited as C. cladosporioides.<br />

DISCUSSION<br />

In the present study, a multilocus genealogy supported by light<br />

<strong>and</strong> SEM microscopy, <strong>and</strong> cultural characteristics was used to<br />

redefine species borders within <strong>Cladosporium</strong>, especially within<br />

the C. herbarum complex. Most of the diagnostic features used<br />

for species delimitation on host material (Heuchert et al. 2005,<br />

Schubert 2005), proved to be applicable in culture. However,<br />

morphological features were often more pronounced in vivo than<br />

in vitro. For instance, conidiophore arrangement is not applicable<br />

to cultures, conidiophore <strong>and</strong> conidium widths were often<br />

narrower in culture than on the natural host, <strong>and</strong> macro- as well<br />

as microconidiophores were often observed in culture, but not on<br />

host material. All species belonging to the C. herbarum complex<br />

are characterised by possessing conidia which are ornamentated,<br />

the ornamentation ranging from minutely verruculose to verrucose,<br />

echinulate or spiny whereas in the C. sphaerospermum complex<br />

species with both smooth-walled as well as ornamented conidia<br />

are included (Zalar et al. 2007). <strong>The</strong> surface ornamentation varies<br />

based on the length of surface protuberances <strong>and</strong> in the density<br />

of ornamentation. Furthermore, the conidia are mainly catenate,<br />

formed in unbranched or branched chains. However, species<br />

previously referred to the <strong>genus</strong> Heterosporium, which usually<br />

produce solitary conidia or unbranched chains of two or three<br />

conidia at the most on the natural host, also belong to this species<br />

complex (e.g., C. iridis). In vitro these chains can become longer<br />

<strong>and</strong> may even be branched. <strong>The</strong> conidiophores formed in culture<br />

are mostly macro- but may also be micronematous, sometimes<br />

forming different types of conidia that vary in shape <strong>and</strong> size from<br />

each other. Most of the species possess nodulose conidiophores<br />

with the conidiogenesis confined to the usually lateral swellings.<br />

However, this phenetic trend is not consistently expressed in all<br />

of the species belonging to the C. herbarum complex. <strong>The</strong> various<br />

<strong>Cladosporium</strong> species within the C. herbarum complex were<br />

observed to have subtle differences in their phenotype which were<br />

visible via cryo-electron microscopy (cryoSEM), <strong>and</strong> are discussed<br />

below.<br />

Fungal colonies: CryoSEM provides the opportunity to study the<br />

organisation of the fungal colony at relatively low magnifications.<br />

<strong>Cladosporium</strong> tenellum proved to be the only fungus able to form<br />

153


Schubert et al.<br />

aerial hyphal str<strong>and</strong>s under the conditions studied. <strong>Cladosporium</strong><br />

variabile formed abundant aerial hyphae, but in C. spinulosum<br />

these were sparse, <strong>and</strong> only conidiophores were observed on the<br />

agar surface. Three-day-old colonies of C. subinflatum formed<br />

numerous, long aerial hyphae, <strong>and</strong> no conidiophores could be<br />

discerned under the binocular. After 11 d the aerial hyphae seemed<br />

to have disappeared, giving rise to conidiophores. <strong>Cladosporium</strong><br />

antarcticum, C. variabile <strong>and</strong> C. ramotenellum showed very large,<br />

swollen (> 10 μm) cells which gave rise to conidiophores. With C.<br />

variabile possible earlier stages of these cells were visible (Fig. 48),<br />

which gave rise to conidiophores. More than one conidiophore could<br />

be formed on such a structure (C. variabile <strong>and</strong> C. ramotenellum).<br />

<strong>Cladosporium</strong> herbarum has very wide hyphae on the agar surface,<br />

which gave rise to conidiophores as lateral branches. <strong>The</strong>se wide<br />

hyphae were observed to anastomose, which may provide a firm<br />

interconnected supporting mycelium for these conidiophores. In<br />

C. herbaroides these wide hyphae could also be discerned, but<br />

conidiophore formation was less obvious. Similarily, C. tenellum<br />

has wide, parallel hyphae that gave rise to conidiophores.<br />

<strong>The</strong>se observations reveal fungal structures in <strong>Cladosporium</strong><br />

that have not previously been reported on, <strong>and</strong> that raise intriguing<br />

biological questions. For instance, why are hyphal str<strong>and</strong>s observed<br />

in some species (C. tenellum), <strong>and</strong> not in others, <strong>and</strong> what happens<br />

to the aerial hyphae during incubation in some species such as C.<br />

subinflatum? Furthermore, these preliminary results suggest that<br />

CryoSEM provide additional features that can be used to distinguish<br />

the different species in the C. herbarum complex.<br />

Fine details of morphological stuctures: CryoSEM provides the<br />

opportunity to study fine details of the conidiophore, (ramo)conidia<br />

<strong>and</strong> scars. Samples can be studied at magnification up to<br />

× 8 000, revealing details at a refinement far above what is<br />

possible under the light microscope (LM) (Fig. 2). However, the LM<br />

micrographs provide information about the different compartments<br />

of ramoconidia, as well as the thickness <strong>and</strong> pigmentation of the<br />

cell wall of different structures. With other words, the different<br />

techniques are complementary, <strong>and</strong> both reveal fungal details that<br />

build up the picture that defines a fungal species.<br />

Conidiophores can vary with respect to their width <strong>and</strong> the<br />

length. <strong>Cladosporium</strong> ramotenellum, C. antarcticum <strong>and</strong> C. variabile<br />

have tapered conidiophores formed on large globoid “foot cells”.<br />

<strong>The</strong> conidiophore itself can be branched. <strong>Cladosporium</strong> spinulosum<br />

has conidiophores that rise from the agar surface, but can have<br />

a common point of origin. <strong>The</strong>se conidiophores are not tapered,<br />

but parallel <strong>and</strong> slender. <strong>The</strong> conidiophores of C. bruhnei <strong>and</strong> C.<br />

herbaroides are rather long, <strong>and</strong> can appear as aerial hyphae.<br />

An important feature of the conidiophore is the location were the<br />

conidia are formed. Conidiophore ends can be simple <strong>and</strong> tubular,<br />

or rounded to more complex, several times geniculate, with several<br />

scars. Conidiophore ends become more elaborate over time.<br />

<strong>Cladosporium</strong> spinulosum <strong>and</strong> C. tenellum have nearly tubular<br />

conidiophore ends, with often very closely aggregated scars. <strong>The</strong><br />

conidiophore ends of C. subinflatum are also near tubular with a<br />

hint of bulbousness. <strong>Cladosporium</strong> subtilissimum is <strong>similar</strong>, but with<br />

somewhat more elevated scars that look denticulate. <strong>Cladosporium</strong><br />

variabile has nodulose, somewhat swollen apices with often sessile,<br />

almost inconspicuous scars. In the case of C. macrocarpum, these<br />

structures are also nodulose to nodose <strong>and</strong> somewhat bent, with<br />

only slightly protuberant loci. <strong>Cladosporium</strong> ramotenellum has<br />

tubular conidiophore ends with pronounced scars. <strong>Cladosporium</strong><br />

antarcticum has very characteristic, tapered ends, <strong>and</strong> widely<br />

dispersed (5 μm) scars. More complex conidiophore ends are<br />

more irregular in shape, <strong>and</strong> have scars dispersed over a longer<br />

distance, such as observed in C. bruhnei, C. herbaroides, <strong>and</strong> C.<br />

herbarum.<br />

Secondary ramoconidia are usually the first conidia formed on<br />

a conidiophore. <strong>The</strong>y are often multicellular, <strong>and</strong> have one basal<br />

cladosporioid hilum, <strong>and</strong> more at the apex. Few <strong>Cladosporium</strong><br />

species additionally form true ramoconidia representing apical<br />

parts of the conidiophore which secede at a septum resulting in<br />

an undifferentiated non-coronate base <strong>and</strong> function as conidia.<br />

Ramification of conidial chains is realised through these conidia.<br />

<strong>The</strong>y can occur in up to three stages, which results in elaborated<br />

spore structures. <strong>The</strong> basal secondary ramoconidium is invariably<br />

the largest, <strong>and</strong> cell size decreases through a series of additional<br />

secondary ramoconidia, intercalary conidia, <strong>and</strong> small, terminal<br />

conidia. <strong>The</strong> elongation of secondary ramoconidia varies among<br />

the different species. <strong>Cladosporium</strong> macrocarpum has broadly<br />

ellipsoid to cylindrical secondary ramoconidia usually with broadly<br />

rounded ends, like C. variabile, while C. spinulosum has secondary<br />

ramoconidia that can often hardly be discerned from the conidia<br />

that are formed at later stages. <strong>The</strong> conidia of the other species<br />

roughly fall between these species. <strong>The</strong> most notable structures on<br />

these conidia are their ornamentation, scar pattern <strong>and</strong> morphology.<br />

<strong>Cladosporium</strong> spinulosum forms numerous globose to subsphaerical<br />

spores with digitate, non-tapered surface ornamentation, which<br />

is unique for all the species discussed here. In his study on<br />

<strong>Cladosporium</strong> wall ornamentation, David (1997) recognised three<br />

classes of echinulate surfaces (aculeate, spinulose, digitate), <strong>and</strong><br />

five classes of verrucose surfaces (muricate, granulate, colliculate,<br />

pustulate <strong>and</strong> pedicellate) (Fig. 2). <strong>The</strong> ornamentation particles vary<br />

in shape, width, height <strong>and</strong> density. <strong>The</strong> most strongly ornamented<br />

conidia of the species examined by SEM are formed by C. ossifragi,<br />

with the ornamentation both large (up to 0.5 μm wide) <strong>and</strong> high,<br />

<strong>and</strong> can be regarded as densely muricately ornamented. Strong<br />

ornamentation is also seen in C. herbaroides, which is mostly<br />

granulate. <strong>Cladosporium</strong> tenellum (with muricate, granulate <strong>and</strong><br />

colliculate tendencies) <strong>and</strong> C. bruhnei (mostly granulate with some<br />

muricate projections) have relatively large ornamentation structures<br />

with slightly more space between the units than the other two<br />

species. <strong>Cladosporium</strong> antarcticum, C. ramotenellum, C. variabile<br />

<strong>and</strong> C. subtilissimum exhibit rather large granulate ornamentations<br />

that have a more irregular <strong>and</strong> variable shape. <strong>Cladosporium</strong><br />

subinflatum shows the widest dispersed structures of the series,<br />

being muricate. In contrast, C. macrocarpum has a very neat <strong>and</strong><br />

regular pattern of muricate ornamentation. <strong>The</strong> area of formation<br />

of new spores on conidia is invariably not ornamented, <strong>and</strong> hila all<br />

have the typical <strong>Cladosporium</strong> morphology with a central dome <strong>and</strong><br />

a ring-like structure around it.<br />

Branching patterns: Spores usually show a “line of weakness”<br />

between them where the coronate scars form. It seems that scars at<br />

both sides of the line of weakness have the central dome structure,<br />

which appears to play a major role in the effective mechanism<br />

<strong>Cladosporium</strong> employs for spore dispersal, with the dome actively<br />

pushing the conidia apart. This mechanism is also illustrated in<br />

David (1997, fig. 2E). Indeed, conidia of <strong>Cladosporium</strong> are very<br />

easily dislodged; even snap freezing or the electrical forces inside<br />

the SEM often result in dislodgement of the spores in a powdery<br />

“wave”. It is no surprise, therefore, that <strong>Cladosporium</strong> conidia are<br />

to be found in most air samples. In <strong>Cladosporium</strong>, conidia are<br />

mostly formed in chains, with the size invariably decreasing from<br />

the base to the apex of the row. Upon formation each conidium is<br />

separated from the conidiophore, or previously formed conidium,<br />

<strong>and</strong> hence from its nutrients. <strong>The</strong> basal ramoconidium or secondary<br />

ramoconidia have the nutrients <strong>and</strong> metabolic power to produce<br />

154


<strong>Cladosporium</strong> herbarum species complex<br />

a number of additional secondary ramoconidia that in turn could<br />

produce a chain of intercalary conidia, <strong>and</strong> finally, some small,<br />

single-celled, terminal conidia. Further research is still necessary<br />

to determine if specific branching patterns can be linked to different<br />

species.<br />

A surprising finding from the present study is the huge diversity<br />

in species <strong>and</strong> genotypes that exist in nature, be it in the indoor<br />

environment, on fruit surfaces, or in extreme ecological niches such<br />

as salterns, etc. It is clear that detailed studies would be required<br />

to find <strong>and</strong> characterise other species of <strong>Cladosporium</strong> <strong>and</strong> obtain<br />

a better underst<strong>and</strong>ing of their host ranges <strong>and</strong> ecology. A further<br />

surprise lay in the fact that several of these species are capable<br />

of sexual reproduction, <strong>and</strong> readily form Davidiella teleomorphs in<br />

culture. <strong>The</strong> Davidiella states induced here were all from homothallic<br />

species. Further attention now needs to be given to elucidating<br />

teleomorphs from other species which, as in Mycosphaerella<br />

(Groenewald et al. 2006, Ware et al. 2007) could be heterothallic,<br />

<strong>and</strong> experiencing cl<strong>and</strong>estine sex.<br />

Despite the occurrence of many different genotypes in<br />

variable genes, the degree of diversity in the entire data set was<br />

low. For the majority of the species ITS was almost invariant, with<br />

only six genotypes in the entire dataset. This suggests a very<br />

recent evolution. <strong>The</strong> st<strong>and</strong>ardised index of association (I S A ) was<br />

high (0.3914), indicating an overabundance of clonality <strong>and</strong> / or<br />

inbreeding, the latter possibly matching with observed homothallism<br />

of Davidiella teleomorphs. Clonality was visualised with SplitsTree<br />

software, where star-shaped representations without any sign of<br />

reticulation were obtained for all genes, though at different branch<br />

lengths (Fig. 5). With Structure software an optimal subdivision<br />

was achieved at six putative groups. Some of them were distinctly<br />

separated, yielding a theta (θ) around 0.14, but in most cases there<br />

was considerable overlap in representation of motifs, with θ at<br />

significantly higher values. Results are difficult to interpret due to<br />

the small size of the data set compared to the number of predicted<br />

groups, <strong>and</strong> due to unknown but probably large sampling effects.<br />

With optimal subdivision of the 79 strains at a hypothesised value<br />

of K = 6 (Fig. 4), still a large degree of inter-group <strong>similar</strong>ity was<br />

noted, as was the case at any other level of K. This was particularly<br />

obvious when data from the most variable genes (EF <strong>and</strong> ACT) are<br />

superimposed (Fig. 4). <strong>The</strong> ACT groups are further subdivided by EF<br />

data, but in many cases the same EF motif (indicated with arrows)<br />

was encountered in different (multilocus) species, for example in C.<br />

antarcticum, C. spinulosum, Davidiella sp., <strong>and</strong> the various clusters<br />

comprising <strong>Cladosporium</strong> strains which are phenotypically almost<br />

indistinguishable but genetically distinct from C. subtilissimum. A<br />

<strong>similar</strong> situation was found with the distribution of EF genotypes<br />

(indicated with doughnuts) in C. herbarum <strong>and</strong> C. macrocarpum.<br />

Nevertheless, the data set showed significant structuring, partly<br />

correlating with geography, e.g. the EF-determined cluster of C.<br />

bruhnei that contained isolates from different sources in <strong>The</strong><br />

Netherl<strong>and</strong>s. Differences may be over-accentuated by known<br />

sampling effects, particularly in C. herbarum <strong>and</strong> C. macrocarpum,<br />

where single-spore isolates from a single collection are included.<br />

Taken together the data suggest a recent, preponderantly clonal<br />

evolution, combined with limited natural selection at a low level of<br />

evolutionary pressure. As a result, many genotypes produced by<br />

hot spots in the genes analysed have survived, leading to nearly<br />

r<strong>and</strong>om variation in the data set. Many combinations of motifs that<br />

possibly could emerge have maintained in the course of time due<br />

to the absence of recombination. This indicates that the observed<br />

structure is that of populations within a single species, <strong>and</strong><br />

consequently a distinction of clonal “species” could be redundant.<br />

This conclusion is underlined by the fact that a single source in<br />

a single location can be colonised by various genotypes, such<br />

as grapes in the U.S.A. containing three different, closely related<br />

genotypes. However, the phenomenon of co-inhabitation by<br />

different Mycosphaerella species on the same lesion of Eucalyptus<br />

has been described before (Crous 1998, Crous et al. 2004) <strong>and</strong> it<br />

is therefore not surprising that different genotypes occurring close<br />

together are also observed for the related <strong>genus</strong> <strong>Cladosporium</strong>.<br />

<strong>The</strong>re is no obvious ecological difference between genotypes, <strong>and</strong><br />

hence isolates seem to have equal fitness.<br />

However, in general we noticed a remarkable concordance<br />

of genetic <strong>and</strong> phenetic characters. <strong>The</strong> morphological study was<br />

done prior to sequencing, <strong>and</strong> nearly all morphotypes clustered<br />

in separate molecular entities. <strong>The</strong>re are some exceptions, such<br />

as with C. antarcticum with striking morphology that was almost<br />

identical on the molecular level to <strong>Cladosporium</strong> spp. that resemble<br />

C. subtilissimum <strong>and</strong> would normally have been interpreted to<br />

be a mutant. Conversely, nearly all genetically distinguishable<br />

groups proved to be morphologically different, with the exception<br />

of members of the C. subtilissimum s. lat. complex (indicated as<br />

<strong>Cladosporium</strong> sp. in Fig. 3 <strong>and</strong> Table 1). <strong>The</strong> possibility remains<br />

that the found genetic parameters correlate with phenetic markers<br />

other than morphology, such as virulence, toxins or antifungal<br />

susceptibilities. For this reason we introduce the established<br />

entities here as formal species. <strong>The</strong>y can be diagnosed by ACT<br />

sequencing or by phenetic characters provided in the key. For<br />

simple routine purposes, however, they can be seen <strong>and</strong> treated<br />

as the “C. herbarum complex”, based on their close phylogenetic<br />

relationships.<br />

ACKNOWLEDGEMENTS<br />

Several colleagues from different countries provided material <strong>and</strong> valuable cultures<br />

without which this work would not have been possible. In this regard, we thank<br />

particularly Alina G. Greslebin (CIEFAP, Esquel, Chubut, Argentina), Frank Dugan<br />

<strong>and</strong> Lindsey DuToit (Washington State Univ., U.S.A.), <strong>and</strong> Annette Ramaley<br />

(Colorado, U.S.A.). Konstanze Schubert was financially supported by a Synthesys<br />

grant (No. 2559) which is gratefully acknowledged. We thank Marjan Vermaas for<br />

preparing the photographic plates, Rianne van Dijken for the growth studies, <strong>and</strong><br />

Arien van Iperen for maintaining the cultures studied. Bert Gerrits van den Ende is<br />

acknowledged for assisting with the molecular data analyses.<br />

REFERENCES<br />

Aptroot A (2006). Mycosphaerella <strong>and</strong> its anamorphs: 2. Conspectus of<br />

Mycosphaerella. <strong>CBS</strong> Biodiversity Series 5: 1–231.<br />

Arx JA von (1950). Über die Ascusform von <strong>Cladosporium</strong> herbarum (Pers.) Link.<br />

Sydowia 4: 320–324.<br />

Barr ME (1958). Life history studies of Mycosphaerella tassiana <strong>and</strong> M. typhae.<br />

Mycologia 50: 501–513.<br />

Braun U (2000). Miscellaneous notes on some micromycetes. Schlechtendalia 5:<br />

31–56.<br />

Braun U, Crous PW, Dugan FM, Groenewald JZ, Hoog GS de (2003). Phylogeny<br />

<strong>and</strong> taxonomy of cladosporium-like hyphomycetes, including Davidiella gen.<br />

nov., the teleo-morph of <strong>Cladosporium</strong> s.str. Mycological Progress 2: 3–18.<br />

Braun U, Hill CF, Schubert K (2006). New species <strong>and</strong> new records of biotrophic<br />

micromycetes from Australia, Fiji, New Zeal<strong>and</strong> <strong>and</strong> Thail<strong>and</strong>. Fungal Diversity<br />

22: 13–35.<br />

Brown KB, Hyde KD, Guest DI (1998). Preliminary studies on endophytic fungal<br />

communities of Musa acuminata species complex in Hong Kong <strong>and</strong> Australia.<br />

Fungal Diversity 1: 27–51.<br />

Clements FE, Shear CL (1931). <strong>The</strong> genera of fungi. New York: H.W. Wilson Co.<br />

Corda ACJ (1837). Icones fungorum hucusque cognitorum. Vol. 1. Praha.<br />

Crous PW (1998). Mycosphaerella spp. <strong>and</strong> their anamorphs associated with leaf<br />

spot diseases of Eucalyptus. Mycologia Memoir 21: 1–170.<br />

www.studiesinmycology.org<br />

155


Schubert et al.<br />

Crous PW, Aptroot A, Kang JC, Braun U, Wingfield MJ (2000). <strong>The</strong> <strong>genus</strong><br />

Mycosphaerella <strong>and</strong> its anamorphs. Studies in Mycology 45: 107–121.<br />

Crous PW, Braun U, Schubert K, Groenewald JZ (2007a). Delimiting <strong>Cladosporium</strong><br />

from morphologically <strong>similar</strong> genera. Studies in Mycology 58: 33–56.<br />

Crous PW, Groenewald JZ, Groenewald M, Caldwell P, Braun U, Harrington TC<br />

(2006). Species of Cercospora associated with grey leaf spot of maize. Studies<br />

in Mycology 55: 189–197.<br />

Crous PW, Groenewald JZ, Mansilla JP, Hunter GC, Wingfield MJ (2004).<br />

Phylogenetic reassessment of Mycosphaerella spp. <strong>and</strong> their anamorphs<br />

occurring on Eucalyptus. Studies in Mycology 50: 195–214.<br />

Crous PW, Kang JC, Braun U (2001). A phylogenetic redefinition of anamorph<br />

genera in Mycosphaerella based on ITS rDNA sequences. Mycologia 93:<br />

1081–1101.<br />

David JC (1995a). <strong>Cladosporium</strong> magnusianum. Mycopathologia 129: 53–54.<br />

David JC (1995b). <strong>Cladosporium</strong> varibile. Mycopathologia 129: 57–58.<br />

David JC (1997). A contribution to the systematics of <strong>Cladosporium</strong>. Revision of the<br />

fungi previously referred to Heterosporium. Mycological Papers 172: 1–157.<br />

Domsch KH, Gams W, Anderson TH (1980). Compendium of soil fungi. Vols 1 & 2.<br />

Acad. Press, London.<br />

Dugan FM, Roberts RG (1994). Morphological <strong>and</strong> reproductive aspects of<br />

<strong>Cladosporium</strong> macrocarpum <strong>and</strong> C. herbarum from bing cherry fruits.<br />

Mycotaxon 52: 513–522.<br />

Dugan FM, Schubert K, Braun U (2004). Check-list of <strong>Cladosporium</strong> names.<br />

Schlechtendalia 11: 1–103.<br />

El-Morsy EM (2000). Fungi isolated from the endorhizosphere of halophytic plants<br />

from the Red Sea Coast of Egypt. Fungal Diversity 5: 43–54.<br />

Ellis MB (1971). Dematiaceous hyphomycetes. Commonwealth Mycological<br />

Institute, Kew.<br />

Ellis MB (1972). Dematiaceous hyphomycetes. XI. Mycological Papers 131: 1–25.<br />

Ellis MB, Ellis JP (1985). Microfungi on l<strong>and</strong> plants. An identification h<strong>and</strong>book.<br />

MacMillan, New York.<br />

Ellis MB, Ellis JP (1988). Microfungi on miscellaneous substrates. An identification<br />

h<strong>and</strong>book. Croom Helm, London, <strong>and</strong> Timber Press, Portl<strong>and</strong>, Oregon.<br />

Ellis MB, Waller JM (1974). Mycosphaerella macrospora (conidial state: <strong>Cladosporium</strong><br />

iridis). CMI Descriptions of Pathogenic Fungi <strong>and</strong> Bacteria No. 435.<br />

Falush D, Stephens M, Pritchard JK (2003). Inference of population structure:<br />

Extensions to linked loci <strong>and</strong> correlated allele frequencies. Genetics 164:<br />

1567–1587.<br />

Gams W, Verkleij GJM, Crous PW (2007). <strong>CBS</strong> Course of Mycology, 5 th ed.<br />

Centraalbureau voor Schimmelcultures, Utrecht, Netherl<strong>and</strong>s.<br />

Groenewald M, Groenewald JZ, Harrington TC, Abeln ECA, Crous PW (2006).<br />

Mating type gene analysis in apparently asexual Cercospora species is<br />

suggestive of cryptic sex. Fungal Genetics <strong>and</strong> Biology 43: 813–825.<br />

Haubold B, Hudson RR (2000). LIAN 3.0: detecting linkage disequilibrium in<br />

multilocus data. Bioinformatics Applications Note 16: 847–848.<br />

Haubold B, Travisano M, Raibey PB, Hudson RR (1998). Detecting linkage<br />

disequilibrium in bacterial populations. Genetics 50: 1341–1348.<br />

Hawksworth DL (1979). <strong>The</strong> lichenicolous hyphomycetes. Bulletin of the British<br />

Museum (Natural History), Botany 6: 183–300.<br />

Heuchert B, Braun U (2006). On some <strong>dematiaceous</strong> lichenicolous hyphomycetes.<br />

Herzogia 19: 11–21.<br />

Heuchert B, Braun U, Schubert K (2005). Morphotaxonomic revision of fungicolous<br />

<strong>Cladosporium</strong> species (hyphomycetes). Schlechtendalia 13: 1–78.<br />

Ho MHM, Castañeda RF, Dugan FM, Jong SC (1999). <strong>Cladosporium</strong> <strong>and</strong><br />

Cladophialophora in culture: descriptions <strong>and</strong> an exp<strong>and</strong>ed key. Mycotaxon<br />

72: 115–157.<br />

Hoog GS de, Guarro J, Gené J, Figueras MJ (2000). Atlas of clinical fungi, 2nd ed.<br />

Centraalbureau voor Schimmelcultures, Utrecht <strong>and</strong> Universitat rovira I virgili,<br />

Reus.<br />

Hughes SJ (1958). Revisiones hyphomycetum aliquot cum appendice de nominibus<br />

rejiciendis. Canadian Journal of Botany 36: 727–836.<br />

Huson DH, Bryant D (2006). Application of phylogenetic networks in evolutionary<br />

studies. Molecular Biology <strong>and</strong> Evolution 23: 254–267.<br />

Jaap O (1902). Abh<strong>and</strong>lungen. Schriften des Naturwissenschaftlichen Vereins<br />

Schleswig-Holstein 12: 346–347.<br />

Jolley KA, Feil EJ, Chan MS, Maiden MCJ (2001). Sequence type analysis <strong>and</strong><br />

recombinational tests (START). Bioinformatics Applications Notes 17: 1230–<br />

1231.<br />

Kirk PM, Cannon PF, David JC, Stalpers JA (2001). Dictionary of the fungi, 9 th edn.<br />

CABI Biosciences, Egham, U.K.<br />

Lind J (1913). Danish fungi as represented in the herbarium of E. Rostrup.<br />

Copenhagen.<br />

Linder DH (1947). Botany of the Canadian Eastern Arctic, part II Thallophyta <strong>and</strong><br />

Bryophyta. 4. Fungi. Bulletin of the National Museum of Canada 97, Biological<br />

Series 26: 234–297.<br />

Link HF (1809). Observationes in ordines plantarum naturales. Dissertatio I,<br />

complectens An<strong>and</strong>rarum ordines Epiphytas, Mucedines, Gastromycos et<br />

Fungos. Der Gesellschaft naturforschender Freunde zu Berlin Magazin für die<br />

neuesten Entdeckungen in der gesammten Naturkunde 3: 3–42.<br />

Link HF (1816). Observationes in ordines plantarum naturales. Dissertatio II, sistens<br />

nuperas de Mucedinum et Gastromycorum ordinibus observationes. Der<br />

Gesellschaft naturforschender Freunde zu Berlin Magazin für die neuesten<br />

Entdeckungen in der gesammten Naturkunde 7: 25–45.<br />

Matsushima T (1985). Matsushima Mycological Memoirs No. 4. Kobe.<br />

McKemy JM, Morgan-Jones G (1990). Studies in the <strong>genus</strong> <strong>Cladosporium</strong> sensu<br />

lato II. Concerning Heterosporium gracile, the causal organism of leaf spot<br />

disease of Iris species. Mycotaxon 39: 425–440.<br />

McKemy JM, Morgan-Jones G (1991). Studies in the <strong>genus</strong> <strong>Cladosporium</strong> sensu<br />

lato V. Concerning the type species <strong>Cladosporium</strong> herbarum. Mycotaxon 42:<br />

307–317.<br />

Nyl<strong>and</strong>er JAA (2004). MrAic.pl. Program distributed by the author (http://www.abc.<br />

se/~nyl<strong>and</strong>er/). Evolutionary Biology Centre, Uppsala University.<br />

Persoon CH (1794). Neuer Versuch einer systematischen Eintheilung der<br />

Schwämme. Neues Magazin für die Botanik in ihrem ganzen Umfange 1:<br />

63–128.<br />

Persoon CH (1822). Mycologia europaea. Sectio prima. Erlangen.<br />

Prasil K, Hoog GS de (1988). Variability in <strong>Cladosporium</strong> herbarum. Transactions of<br />

the British Mycological Society 90: 49–54.<br />

Pritchard JK, Stephens M, Donnelly P (2000). Inference of population structure<br />

using multilocus genotype data. Genetics 155: 945–959.<br />

Rayner RW (1970). A mycological colour chart. CMI <strong>and</strong> British Mycological Society.<br />

Kew, Surrey, Engl<strong>and</strong>.<br />

Riesen T, Sieber T (1985). Endophytic fungi in winter wheat (Triticum aestivum L.).<br />

Swiss Federal Institute of Technology, Zürich.<br />

Saccardo PA (1899). Sylloge Fungorum vol. 14 (Saccardo, P.A. & Sydow, P. eds.).<br />

Padova.<br />

Samson RA, Houbraken JAMP, Summerbell RC, Flannigan B, Miller JD (2001).<br />

Common <strong>and</strong> important species of Actinomycetes <strong>and</strong> fungi in indoor<br />

environments. In: Microorganisms in Home <strong>and</strong> Indoor Work Environments:<br />

Diversity, Health Impacts, Investigation <strong>and</strong> Control (Flannigan B, Samson RA,<br />

Miller JD, eds). Taylor & Francis, London: 287–473.<br />

Samson RA, Reenen-Hoekstra, ES van, Frisvad, JC, Filtenborg O (2000).<br />

Introduction to food- <strong>and</strong> airborne fungi, 6 th ed. Centraalbureau voor<br />

Schimmelcultures, Utrecht.<br />

Schoch C, Shoemaker RA, Seifert KA, Hambleton S, Spatafora JW, Crous PW<br />

(2006). A multigene phylogeny of the Dothideomycetes using four nuclear loci.<br />

Mycologia 98: 1041–1052.<br />

Schubert K (2005). Morphotaxonomic revision of foliicolous <strong>Cladosporium</strong> species<br />

(hyphomycetes). Ph.D. dissertation. Martin-Luther-University Halle-Wittenberg,<br />

Germany. http://sundoc.bibliothek.uni-halle.de/diss-online/05/05H208/index.<br />

htm.<br />

Schubert K, Braun U, Mułenko W (2006). Taxonomic revision of the <strong>genus</strong><br />

<strong>Cladosporium</strong> s. lat. 5. Validation <strong>and</strong> description of new species.<br />

Schlechtendalia 14: 55–83.<br />

Shin HD, Lee HT, Im DJ (1999). Occurence of German Iris leaf spot caused by<br />

<strong>Cladosporium</strong> iridis in Korea. Plant Pathology Journal 15: 124–126.<br />

Sivanesan A (1984). <strong>The</strong> bitunicate Ascomycetes <strong>and</strong> their anamorphs. Cramer<br />

Verlag, Vaduz.<br />

Vries GA de (1952). Contribution to the knowledge of the <strong>genus</strong> <strong>Cladosporium</strong> Link<br />

ex Fr. <strong>CBS</strong>, Baarn.<br />

Wang CJK, Zabel RA (1990). Identification manual for fungi from utility poles in the<br />

eastern United States. ATCC, Rockville, MD.<br />

Ware SB, Verstappen ECP, Breeden J, Cavaletto JR, Goodwin SB, Waalwijk C,<br />

Crous PW, Kema GHJ (2007). Discovery of a functional Mycosphaerella<br />

teleomorph in the presumed asexual barley pathogen Septoria passerinii.<br />

Fungal Genetics <strong>and</strong> Biology 44: 389–397.<br />

Zalar P, Hoog GS de, Schroers HJ, Crous PW, Groenewald JZ & Gunde-Cimerman<br />

N (2007). Phylogeny <strong>and</strong> ecology of the ubiquitous saprobe <strong>Cladosporium</strong><br />

sphaerospermum, with description of seven new species from hypersaline<br />

environments. Studies in Mycology 58: 157–183.<br />

156


available online at www.studiesinmycology.org<br />

doi:10.3114/sim.2007.58.06<br />

Studies in Mycology 58: 157–183. 2007.<br />

Phylogeny <strong>and</strong> ecology of the ubiquitous saprobe <strong>Cladosporium</strong> sphaerospermum,<br />

with descriptions of seven new species from hypersaline environments<br />

P. Zalar 1* , G.S. de Hoog 2,3 , H.-J. Schroers 4 , P.W. Crous 2 , J.Z. Groenewald 2 <strong>and</strong> N. Gunde-Cimerman 1<br />

1<br />

Biotechnical Faculty, Department of Biology, Večna pot 111, SI-1000 Ljubljana, Slovenia; 2 <strong>CBS</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD Utrecht, <strong>The</strong> Netherl<strong>and</strong>s;<br />

3<br />

Institute for Biodiversity <strong>and</strong> Ecosystem Dynamics, University of Amsterdam, Kruislaan 315, 1098 SM Amsterdam, <strong>The</strong> Netherl<strong>and</strong>s; 4 Agricultural Institute of Slovenia,<br />

Hacquetova 17, p.p. 2553, 1001 Ljubljana, Slovenia<br />

*Correspondence: Polona Zalar, polona.zalar@bf.uni-lj.si<br />

Abstract: Saprobic <strong>Cladosporium</strong> isolates morphologically <strong>similar</strong> to C. sphaerospermum are phylogenetically analysed on the basis of DNA sequences of the ribosomal RNA<br />

gene cluster, including the internal transcribed spacer regions ITS1 <strong>and</strong> ITS2, the 5.8S rDNA (ITS) <strong>and</strong> the small subunit (SSU) rDNA as well as β-tubulin <strong>and</strong> actin gene introns<br />

<strong>and</strong> exons. Most of the C. sphaerospermum-like species show halotolerance as a recurrent feature. <strong>Cladosporium</strong> sphaerospermum, which is characterised by almost globose<br />

conidia, is redefined on the basis of its ex-neotype culture. <strong>Cladosporium</strong> dominicanum, C. psychrotolerans, C. velox, C. spinulosum <strong>and</strong> C. halotolerans, all with globoid conidia,<br />

are newly described on the basis of phylogenetic analyses <strong>and</strong> cryptic morphological <strong>and</strong> physiological characters. <strong>Cladosporium</strong> halotolerans was isolated from hypersaline<br />

water <strong>and</strong> bathrooms <strong>and</strong> detected once on dolphin skin. <strong>Cladosporium</strong> dominicanum <strong>and</strong> C. velox were isolated from plant material <strong>and</strong> hypersaline water. <strong>Cladosporium</strong><br />

psychrotolerans, which grows well at 4 °C but not at 30 °C, <strong>and</strong> C. spinulosum, having conspicuously ornamented conidia with long digitate projections, are currently only known<br />

from hypersaline water. We also newly describe C. salinae from hypersaline water <strong>and</strong> C. fusiforme from hypersaline water <strong>and</strong> animal feed. Both species have ovoid to ellipsoid<br />

conidia <strong>and</strong> are therefore reminiscent of C. herbarum. <strong>Cladosporium</strong> langeronii (= Hormodendrum langeronii) previously described as a pathogen on human skin, is halotolerant<br />

but has not yet been recorded from hypersaline environments.<br />

Taxonomic novelties: <strong>Cladosporium</strong> dominicanum Zalar, de Hoog & Gunde-Cimerman, sp. nov., C. fusiforme Zalar, de Hoog & Gunde-Cimerman, sp. nov., C. halotolerans<br />

Zalar, de Hoog & Gunde-Cimerman, sp. nov., C. psychrotolerans Zalar, de Hoog & Gunde-Cimerman, sp. nov., C. salinae Zalar, de Hoog & Gunde-Cimerman, sp. nov.,<br />

C. spinulosum Zalar, de Hoog & Gunde-Cimerman, sp. nov., C. velox Zalar, de Hoog & Gunde-Cimerman, sp. nov.<br />

Key words: Actin, β-tubulin, halotolerance, ITS rDNA, phylogeny, SSU rDNA, taxonomy.<br />

INTRODUCTION<br />

<strong>The</strong> halophilic <strong>and</strong> halotolerant mycobiota from hypersaline<br />

aqueous habitats worldwide frequently contain <strong>Cladosporium</strong><br />

Link isolates (Gunde-Cimerman et al. 2000, Butinar et al. 2005).<br />

Initially, they were considered as airborne contaminants, but<br />

surprisingly, many of these <strong>Cladosporium</strong> isolates were identified<br />

as C. sphaerospermum Penz. because they formed globoid conidia<br />

(data unpublished). <strong>Cladosporium</strong> sphaerospermum, known as<br />

one of the most common air-borne, cosmopolitan <strong>Cladosporium</strong><br />

species, was frequently isolated from indoor <strong>and</strong> outdoor air (Park<br />

et al. 2004), dwellings (Aihara et al. 2001), <strong>and</strong> occasionally from<br />

humans (Badillet et al. 1982) <strong>and</strong> plants (Pereira et al. 2002).<br />

Strains morphologically identified as C. sphaerospermum were<br />

able to grow at a very low water activity (a w<br />

0.816), while other<br />

cladosporia clearly preferred a higher, less extreme water activity<br />

(Hocking et al. 1994). This pronounced osmotolerance suggests<br />

a predilection for osmotically stressed environments although<br />

C. sphaerospermum is reported from a wide range of habitats<br />

including osmotically non-stressed niches.<br />

We therefore hypothesised that C. sphaerospermum<br />

represents a complex of species having either narrow or wide<br />

ecological amplitudes. <strong>The</strong> molecular diversity of strains identified<br />

as C. sphaerospermum has not yet been determined <strong>and</strong> isolates<br />

from humans have not yet been critically compared with those from<br />

environmental samples. <strong>The</strong>refore, a taxonomic study was initiated<br />

with the aim to define phylogenetically <strong>and</strong> morphologically distinct<br />

entities <strong>and</strong> to describe their in vitro osmotolerance <strong>and</strong> their natural<br />

ecological preferences.<br />

MATERIALS AND METHODS<br />

Sampling<br />

Samples of hypersaline water were collected from salterns located<br />

at different sites of the Mediterranean basin (Slovenia, Bosnia <strong>and</strong><br />

Herzegovina, Spain), different coastal areas along the Atlantic<br />

Ocean (Monte Cristy, Dominican Republic; Swakopmund, Namibia),<br />

the Red Sea (Eilat, Israel), the Dead Sea (Ein Gedi, Israel), <strong>and</strong><br />

the salt Lake Enriquillio (Dominican Republic). Samples from the<br />

Sečovlje salterns (Slovenia) were collected once per month in 1999.<br />

Samples from the Santa Pola salterns <strong>and</strong> Ebre delta river saltern<br />

(Spain) were taken twice (July <strong>and</strong> November) in 2000. A saltern in<br />

Namibia <strong>and</strong> one in the Dominican Republic were sampled twice<br />

(August <strong>and</strong> October) in 2002. Various salinities, ranging from 15 to<br />

32 % NaCl were encountered in these ponds.<br />

Isolation <strong>and</strong> maintenance of fungi<br />

Strains were isolated from salterns using filtration of hypersaline<br />

water through membrane filters (pore diam 0.45 μm), followed by<br />

incubation of the membrane filters on different culture media with<br />

lowered water activity (Gunde-Cimerman et al. 2000). Only colonies<br />

of different morphology on one particular selective medium per<br />

sample were analysed further. Strains were carefully selected from<br />

different evaporation ponds, collected at different times, in order to<br />

avoid sampling of identical clones. Subcultures were maintained<br />

at the Culture Collection of Extremophilic Fungi (EXF, Biotechnical<br />

Faculty, Ljubljana, Slovenia), while a selection was deposited at<br />

the Centraalbureau voor Schimmelcultures (<strong>CBS</strong>, Utrecht, <strong>The</strong><br />

Netherl<strong>and</strong>s) <strong>and</strong> the Culture Collection of the National Institute<br />

of Chemistry (MZKI, Ljubljana, Slovenia). Reference strains were<br />

obtained from <strong>CBS</strong>, <strong>and</strong> were selected either on the basis of the<br />

strain history, name, or on the basis of their ITS rDNA sequence.<br />

Strains were maintained on oatmeal agar (OA; diluted OA, Difco:<br />

15 g of Difco 255210 OA medium, 12 g of agar, dissolved in 1 L<br />

of distilled water) with or without 5 % additional NaCl. <strong>The</strong>y were<br />

preserved in liquid nitrogen or by lyophilisation. Strains studied are<br />

listed in Table 1.<br />

157


Zalar et al.<br />

Table 1. List of <strong>Cladosporium</strong> strains, with their current <strong>and</strong> original names, geography, GenBank accession numbers <strong>and</strong> references to earlier published sequences.<br />

Strain Nr. a Source Geography GenBank accession Nr. b<br />

ITS rDNA / 18S rDNA actin β-tubulin<br />

<strong>Cladosporium</strong> bruhnei<br />

<strong>CBS</strong> 177.71 Thuja tincture <strong>The</strong> Netherl<strong>and</strong>s, Amsterdam DQ780399 / DQ780938 EF101354 EF101451<br />

<strong>CBS</strong> 812.71 Polygonatum odoratum, leaf Czech Republic, Lisen DQ780401 / – – –<br />

<strong>Cladosporium</strong> cladosporioides<br />

<strong>CBS</strong> 170.54 NT Arundo, leaf U.K., Engl<strong>and</strong>, Kew AY213640 / DQ780940 EF101352 EF101453<br />

EXF-321 Hypersaline water Slovenia, Sečovlje saltern DQ780408 / – – –<br />

EXF-780 DQ780409 / – – –<br />

EXF-946 Hypersaline water Bosnia <strong>and</strong> Herzegovina, DQ780410 / – – –<br />

Ston saltern<br />

<strong>Cladosporium</strong> dominicanum<br />

CPC 11683 Citrus fruit (orange) Iran DQ780357 / – EF101369 EF101419<br />

EXF-696 Hypersaline water Dominican Republic, saltern DQ780358 / – EF101367 EF101420<br />

EXF-718 Hypersaline water Dominican Republic, salt lake DQ780356 / – EF101370 EF101418<br />

Enriquilio<br />

EXF-720 Hypersaline water Dominican Republic, saltern DQ780355 / – – EF101417<br />

EXF-727 Hypersaline water Dominican Republic, saltern DQ780354 / – – EF101416<br />

EXF-732 T; <strong>CBS</strong> 119415 Hypersaline water Dominican Republic, salt lake DQ780353 / – EF101368 EF101415<br />

Enriquilio<br />

<strong>Cladosporium</strong> fusiforme<br />

<strong>CBS</strong> 452.71 Chicken food Canada DQ780390 / – EF101371 EF101447<br />

EXF-397 Hypersaline water Slovenia, Sečovlje saltern DQ780389 / – EF101373 EF101445<br />

EXF-449 T; <strong>CBS</strong> 119414 Hypersaline water Slovenia, Sečovlje saltern DQ780388 / DQ780935 EF101372 EF101446<br />

<strong>Cladosporium</strong> herbarum<br />

ATCC 66670, as Davidiella tassiana CCA-treated Douglas-fir pole U.S.A., New York, Geneva AY361959 2 & DQ780400 / DQ780939 AY752193 11 EF101452<br />

<strong>Cladosporium</strong> halotolerans<br />

ATCC 26362 Liver <strong>and</strong> intestine of diseased frog U.S.A., New Jersey AY361982 2 / – – –<br />

ATCC 64726 Peanut cell suspension tissue culture U.S.A., Georgia AY361968 2 / – – –<br />

<strong>CBS</strong> 280.49<br />

Stem of Hypericum perforatum Switzerl<strong>and</strong>, Glarus,<br />

DQ780369 / – EF101402 EF101432<br />

identified as Mycosphaerella hyperici Mühlehorn<br />

<strong>CBS</strong> 191.54 Laboratory air Great Britain – / – – –<br />

<strong>CBS</strong> 573.78 Aureobasidium caulivorum Russia, Moscow region – / – – –<br />

<strong>CBS</strong> 626.82 – Sweden, Stockholm – / – – –<br />

dH 12862; EXF-2533 Culture contaminant Brazil DQ780371 / – EF101400 EF101422<br />

dH 12941; EXF-2534 Culture contaminant Turkey – / – EF101421<br />

dH 12991; EXF-2535 Brain Turkey DQ780372 / – EF101423<br />

dH 13911; EXF-2422 Ice Arctics DQ780370 / – EF101401 EF101430<br />

EXF-228; MZKI B-840 Hypersaline water Slovenia, Sečovlje saltern DQ780365 / DQ780930 EF101393 EF101425<br />

EXF-380 Hypersaline water Slovenia, Sečovlje saltern DQ780368 / – EF101394 EF101427<br />

EXF-564 Hypersaline water Namibia, saltern DQ780363 / – EF101395 EF101433<br />

EXF-565 Hypersaline water Namibia, saltern – / – – –<br />

EXF-567 Hypersaline water Namibia, saltern – / – – –<br />

EXF-571 Hypersaline water Namibia, saltern – / – – –<br />

EXF-572 T; <strong>CBS</strong> 119416 Hypersaline water Namibia, saltern DQ780364 / – EF101397 EF101424<br />

EXF-646 Hypersaline water Spain, Santa Pola saltern DQ780366 / – EF101398 EF101428<br />

EXF-698 Hypersaline water Dominican Republic, saltern – / – – –<br />

EXF-703 Hypersaline water Dominican Republic, salt lake DQ780367 / – EF101392 EF101426<br />

Enriquilio<br />

EXF-944 Hypersaline water Bosnia <strong>and</strong> Herzegovina, – / – – –<br />

Ston saltern<br />

EXF-972 Bathroom Slovenia – / – – –<br />

EXF-977 Bathroom Slovenia DQ780362 / – EF101396 EF101431<br />

158


<strong>Cladosporium</strong> sphaerospermum species complex<br />

Strain Nr. a Source Geography GenBank accession Nr. b<br />

ITS rDNA / 18S rDNA actin β-tubulin<br />

EXF-1072 Hypersaline water Israel, Dead Sea DQ780373 / – EF101399 EF101428<br />

EXF-2372 Hypersaline water Slovenia, Sečovlje saltern – / – – –<br />

UAMH 7686<br />

Indoor air ex RCS strip, from Apis<br />

mellifera overwintering facility<br />

U.S.A., Alta, Clyde Corner AY625063 5 / – – –<br />

– rDNA from bottlenose dolphin skin U.S.A., Texas AF035674 6 / – – –<br />

infected with Loboa loboi<br />

– Microcolony, on rock Turkey, Antalya AJ971409 7 / – – –<br />

– Microcolony, on rock Turkey, Antalya AJ971408 7 / – – –<br />

– Tomato leaves – L25433 8 / – – –<br />

<strong>Cladosporium</strong> langeronii<br />

<strong>CBS</strong> 189.54 NT Mycosis Brazil DQ780379 / DQ780932 EF101357 EF101435<br />

<strong>CBS</strong> 601.84 Picea abies, wood Germany, Göttingen DQ780382 / – EF101360 EF101438<br />

<strong>CBS</strong> 101880 Moist aluminium school window frame Belgium, Lichtervoorde DQ780380 / – EF101359 EF101440<br />

<strong>CBS</strong> 109868 Mortar of Muro Farnesiano Italy, Parma DQ780377 / – EF101362 EF101434<br />

dH 11736 Biomat in a lake Antarctics DQ780381 / – EF101363 EF101436<br />

dH 12459 Orig. face lesion Brazil DQ780378 / – EF101358 EF101439<br />

dH 13833 Ice Arctics DQ780383 / – EF101361 EF101437<br />

– Nasal mucus – AF455525 4 / – – –<br />

– Nasal mucus – AY345352 4 / – – –<br />

– Mycorrhizal roots – DQ068982 9 / – – –<br />

<strong>Cladosporium</strong> oxysporum<br />

ATCC 66669 Creosote-treated southern pine pole U.S.A., New York,<br />

AF393689 10 / DQ780395 AY752192 11 EF101454<br />

Binghamton<br />

ATCC 76499 Decayed leaf, Lespedeza bicolor – AF393720 – –<br />

<strong>CBS</strong> 125.80 Cirsium vulgare, seedcoat <strong>The</strong> Netherl<strong>and</strong>s AJ300332 12 / DQ780941 EF101351 EF101455<br />

EXF-697 Hypersaline water Dominican Republic, salt lake DQ780392 / – – –<br />

Enriquilio<br />

EXF-699 Hypersaline water Dominican Republic, saltern DQ780394 / – – –<br />

EXF-710 Hypersaline water Dominican Republic, saltern DQ780393 / – – –<br />

EXF-711 Hypersaline water Dominican Republic, saltern DQ780391 / – – –<br />

<strong>Cladosporium</strong> psychrotolerans<br />

EXF-326 Hypersaline water Slovenia, Sečovlje saltern DQ780387 / DQ780934 – EF101444<br />

EXF-332 Hypersaline water Slovenia, Sečovlje saltern DQ780385 / DQ780933 EF101364 EF101441<br />

EXF-391 T; <strong>CBS</strong> 119412 Hypersaline water Slovenia, Sečovlje saltern DQ780386 / – EF101365 EF101442<br />

EXF-714 Hypersaline water Dominican Republic DQ780384 / – EF101366 EF101443<br />

<strong>Cladosporium</strong> ramotenellum<br />

EXF-454 T; CPC 12043 Hypersaline water Slovenia, Sečovlje saltern DQ780403 / – – –<br />

<strong>Cladosporium</strong> salinae<br />

EXF-322 Hypersaline water Slovenia, Sečovlje DQ780375 / – EF101391 EF101403<br />

EXF-335 T; <strong>CBS</strong> 119413 Hypersaline water Slovenia, Sečovlje DQ780374 / DQ780931 EF101390 EF101405<br />

EXF-604 Hypersaline water Spain, Santa Pola DQ780376 / – EF101389 EF101404<br />

<strong>Cladosporium</strong> sp.<br />

<strong>CBS</strong> 300.96 Soil along coral reef coast Papua New Guinea, Madang, DQ780352 / – EF101385 –<br />

Jais Aben<br />

EXF-595 Hypersaline water Spain, Santa Pola saltern DQ780402 / – – –<br />

<strong>Cladosporium</strong> sphaerospermum<br />

ATCC 12092 Soil Canada AY361988 2 / – – –<br />

ATCC 200384 Compost biofilter <strong>The</strong> Netherl<strong>and</strong>s AY361991 2 / – – –<br />

<strong>CBS</strong> 109.14; ATCC 36950 Carya illinoensis leaf scale U.S.A. DQ780350 / – EF101384 EF101410<br />

<strong>CBS</strong> 122.47; IFO 6377; IMI 49640; Decaying stem of Begonia sp., <strong>The</strong> Netherl<strong>and</strong>s, Aalsmeer AJ244228 1 / – – –<br />

VKM F-772; ATCC 11292<br />

with Thielaviopsis basicola<br />

<strong>CBS</strong> 188.54; ATCC 11290; IMI de Vries (Engelhardt strain) – AY361990 2 & AY251077 3 / – – –<br />

049638<br />

<strong>CBS</strong> 190.54; ATCC 11293; IFO 6380; – – AY361992 2 / – – –<br />

IMI 49641<br />

<strong>CBS</strong> 192.54; ATCC 11288; IMI 49636 Nail of man – AY361989 2 / – – –<br />

www.studiesinmycology.org<br />

159


Zalar et al.<br />

Table 1. (Continued).<br />

Strain Nr. a Source Geography GenBank accession Nr. b<br />

<strong>CBS</strong> 193.54 NT; ATCC 11289; IMI<br />

49637<br />

ITS rDNA / 18S rDNA actin β-tubulin<br />

Human nails – DQ780343 & AY361958 2 / DQ780925 EF101380 EF101406<br />

<strong>CBS</strong> 122.63 Plywood of Betula sp. Finl<strong>and</strong>, Helsinki – / – – –<br />

<strong>CBS</strong> 102045; EXF-2524; MZKI B- Hypersaline water<br />

Spain, Barcelona, Salines de DQ780351 / – EF101378 EF101411<br />

1066<br />

la Trinitat<br />

<strong>CBS</strong> 114065 Outdoor air Germany, Stuttgart – / – – –<br />

CPC 10944 Gardening peat substrate Russia, Kaliningrad DQ780350 / – – –<br />

EXF-131; MZKI B-1005 Hypersaline water Slovenia, Sečovlje saltern AJ238670 1 / – – –<br />

EXF-328 Hypersaline water Slovenia, Sečovlje saltern – / – – –<br />

EXF-385 Hypersaline water Slovenia, Sečovlje saltern – / – – –<br />

EXF-446 Hypersaline water Slovenia, Sečovlje saltern – / – – –<br />

EXF-455 Hypersaline water Slovenia, Sečovlje saltern DQ780349 / – EF101375 EF101412<br />

EXF-458 Hypersaline water Slovenia, Sečovlje saltern DQ780345 / – EF101374 EF101409<br />

EXF-461 Hypersaline water Slovenia, Sečovlje saltern – / – – –<br />

EXF-464 Hypersaline water Slovenia, Sečovlje saltern – / DQ780927 – –<br />

EXF-465 Hypersaline water Slovenia, Sečovlje saltern – / – – –<br />

EXF-598 Hypersaline water Spain, Santa Pola – / – EF101377 –<br />

EXF-644 Hypersaline water Spain, Santa Pola – / – – –<br />

EXF-645 Hypersaline water Spain, Santa Pola – / – – –<br />

EXF-649 Hypersaline water Spain, Santa Pola – / – – –<br />

EXF-715 Hypersaline water Dominican Republic, saltern – / – – –<br />

EXF-738 Bathroom Slovenia DQ780348 / – EF101383 EF101414<br />

EXF-739 Bathroom Slovenia DQ780344 / – EF101381 EF101407<br />

EXF-781; MZKI B-899 Hypersaline water Slovenia, Sečovlje – / – – –<br />

EXF-788 Hypersaline water Slovenia, Sečovlje – / – – –<br />

EXF-962 Bathroom Slovenia DQ780347 / – EF101382 EF101413<br />

EXF-965 Bathroom Slovenia – / – – –<br />

EXF-1069 Hypersaline water Israel, Eilat saltern – / – EF101376 –<br />

EXF-1061 Hypersaline water Israel, Dead Sea DQ780346 / – EF101379 EF101408<br />

EXF-1726 Hypersaline water Israel, Dead Sea – / – – –<br />

EXF-1732 Hypersaline water Israel, Eilat saltern – / DQ780928 – –<br />

– Bryozoa sp. – AJ557744 / – – –<br />

– Nasal mucus – AF455481 4 / – – –<br />

<strong>Cladosporium</strong> spinulosum<br />

EXF-333 Hypersaline water Slovenia, Sečovlje saltern DQ780404 / – – –<br />

EXF-334 T Hypersaline water Slovenia, Sečovlje saltern DQ780406 / – EF101355 EF101450<br />

EXF-382 Hypersaline water Slovenia, Sečovlje saltern DQ780407 / DQ780936 EF101356 EF101449<br />

<strong>Cladosporium</strong> subinflatum<br />

EXF-343 T; CPC 12041 Hypersaline water Slovenia, Sečovlje saltern DQ780405 / – EF101353 EF101448<br />

<strong>Cladosporium</strong> tenuissimum<br />

ATCC 38027 Soil New Caledonia AF393724 / – – –<br />

EXF-324 Hypersaline water Slovenia, Sečovlje saltern – / DQ780926 – –<br />

EXF-371 Hypersaline water Slovenia, Sečovlje saltern DQ780396 / – – –<br />

EXF-452 Hypersaline water Slovenia, Sečovlje saltern DQ780397 / – – –<br />

EXF-563 Hypersaline water Namibia, saltern DQ780398 / – – –<br />

<strong>Cladosporium</strong> velox<br />

<strong>CBS</strong> 119417 T; CPC 11224 Bamboo sp. India, Charidij DQ780361 / DQ780937 EF101388 EF101456<br />

EXF-466 Hypersaline water Slovenia, Sečovlje saltern DQ780359 / – EF101386 –<br />

EXF-471 Hypersaline water Slovenia, Sečovlje saltern DQ780360 / – EF101387 –<br />

160


<strong>Cladosporium</strong> sphaerospermum species complex<br />

Cultivation <strong>and</strong> microscopy<br />

For growth rate determination <strong>and</strong> phenetic description of colonies,<br />

strains were point inoculated on potato-dextrose agar (PDA, Difco),<br />

OA <strong>and</strong> Blakeslee malt extract agar (MEA, Samson et al. 2002)<br />

<strong>and</strong> incubated at 25 °C for 14 d in darkness. Surface colours were<br />

rated using the colour charts of Kornerup & Wanscher (1967). For<br />

studies of microscopic morphology, strains were grown on synthetic<br />

nutrient agar (SNA, Gams et al. 2007) in slide cultures. SNA blocks<br />

of approximately 1 × 1 cm were cut out aseptically, placed upon<br />

sterile microscope slides, <strong>and</strong> inoculated at the upper four edges<br />

by means of a conidial suspension (Pitt 1979). Inoculated agar<br />

blocks were covered with sterile cover slips <strong>and</strong> incubated in moist<br />

chambers for 7 d at 25 °C in darkness. <strong>The</strong> structure <strong>and</strong> branching<br />

pattern of conidiophores were observed at magnifications × 100,<br />

× 200 <strong>and</strong> × 400 in intact slide cultures under the microscope<br />

without removing the cover slips from the agar blocks. For higher<br />

magnifications (× 400, × 1 000) cover slips were carefully removed<br />

<strong>and</strong> mounted in lactic acid with aniline blue.<br />

Morphological parameters<br />

Morphological terms follow David (1997), Kirk et al. (2001) <strong>and</strong><br />

Schubert et al. (2007 – this volume). Conidiophores in <strong>Cladosporium</strong><br />

are usually ascending <strong>and</strong> sometimes poorly differentiated. Though<br />

the initiation point of conidiophore stipes could sometimes be<br />

determined only approximately, their lengths were in some cases<br />

useful for distinguishing morphologically <strong>similar</strong> species when<br />

observed in slide cultures. <strong>The</strong> branching patterns can be rotationally<br />

symmetric or unilateral. Characters of conidial scars were studied<br />

by light <strong>and</strong> scanning electron microscopy (SEM). Conidial chains<br />

show different branching patterns, determined by the numbers of<br />

conidia in unbranched parts, the nature of ramoconidia as well as<br />

their distribution in conidial chains. Measurements are given as (i)<br />

n 1<br />

–n 2<br />

or (ii) (n 1<br />

–)n 3<br />

–n 4<br />

(–n 2<br />

), with n 1<br />

= minimum value observed; n 2<br />

= maximum value observed; n 3<br />

/n 4<br />

= first/third quartile. For conidia<br />

<strong>and</strong> ramoconidia also average values <strong>and</strong> st<strong>and</strong>ard deviations are<br />

listed. <strong>The</strong> values provided are based on at least 25 measurements<br />

for the conidiophores of each strain, <strong>and</strong> at least 50 measurements<br />

for conidia.<br />

Ecophysiology<br />

To determine the degree of halotolerance, strains were pointinoculated<br />

on MEA without <strong>and</strong> with additional NaCl at concentrations<br />

of 5, 10, 17 <strong>and</strong> 20 % NaCl (w/v) <strong>and</strong> incubated at 25 °C for 14 d.<br />

To determine cardinal temperature requirements for growth, plates<br />

were incubated at 4, 10, 25, 30 <strong>and</strong> 37 °C, <strong>and</strong> colony diameters<br />

measured after 14 d of incubation.<br />

DNA extraction, sequencing <strong>and</strong> analysis<br />

For DNA isolation strains were grown on MEA for 7 d. DNA was<br />

extracted according to Gerrits van den Ende & de Hoog (1999)<br />

by mechanical lysis of approx. 1 cm 2 of mycelium. A fragment of<br />

the rDNA including the Internal Transcribed Spacer region 1, 5.8S<br />

rDNA <strong>and</strong> the ITS 2 (ITS) was amplified using the primers V9G<br />

(de Hoog & Gerrits van den Ende 1998) <strong>and</strong> LS266 (Masclaux<br />

et al. 1995). Sequence reactions were done using primers ITS1<br />

<strong>and</strong> ITS4 (White et al. 1990). For amplification <strong>and</strong> sequencing<br />

of the partial actin gene, primers ACT-512F <strong>and</strong> ACT-783R were<br />

applied according to Carbone & Kohn (1999). For amplification <strong>and</strong><br />

sequencing of the β-tubulin gene primers T1 <strong>and</strong> T22 were used<br />

according to O’Donnell & Cigelnik (1997). A BigDye terminator<br />

cycle sequencing kit (Applied Biosystems, Foster City, CA, U.S.A.)<br />

was used in sequence reactions. Sequences were obtained with<br />

an ABI Prism 3700 DNA Analyzer (Applied Biosystems). <strong>The</strong>y<br />

were assembled <strong>and</strong> edited using SeqMan v. 3.61 (DNAStar,<br />

Inc., Madison, U.S.A.). Sequences downloaded from GenBank<br />

are indicated in the trees by their GenBank accession numbers;<br />

newly generated sequences are indicated by strain numbers<br />

(see also Table 1). Sequences were automatically aligned using<br />

ClustalX v. 1.81 (Jeanmougin et al. 1998). <strong>The</strong> alignments were<br />

adjusted manually using MEGA3 (Kumar et al. 2004). Phylogenetic<br />

relationships of the taxa were estimated from aligned sequences<br />

by the maximum parsimony criterion as implemented in PAUP v.<br />

4.0b10 (Swofford 2003). Data sets of the SSU rDNA, ITS rDNA<br />

<strong>and</strong> the β-tubulin <strong>and</strong> actin genes are analysed separately. Species<br />

of <strong>Cladosporium</strong> s. str. were compared with various taxa of the<br />

Mycosphaerellaceae using SSU rDNA sequences <strong>and</strong> Fusicladium<br />

effusum G. Winter (Venturiaceae) as outgroup. <strong>The</strong> other data sets<br />

focus on <strong>Cladosporium</strong> s. str., using <strong>Cladosporium</strong> salinae Zalar,<br />

de Hoog & Gunde-Cimerman as an outgroup, because this species<br />

was most deviant within <strong>Cladosporium</strong> in the SSU rDNA analysis<br />

(see below). Heuristic searches were performed on all characters,<br />

which were unordered <strong>and</strong> equally weighted. Gaps were treated<br />

as missing characters. Starting tree(s) were obtained via stepwise,<br />

r<strong>and</strong>om, 100 times repeated sequence addition. Other parameters<br />

included a “MaxTrees” setting to 9 000, the tree-bisectionreconnection<br />

as branch-swapping algorithm, <strong>and</strong> the “MulTrees”<br />

option set to active. Branch robustness was tested in the parsimony<br />

analysis by 10 000 search replications, each on bootstrapped data<br />

sets using a fast step-wise addition bootstrap analysis. Bootstrap<br />

values larger than 60 are noted near their respective branches.<br />

Newly generated sequences were deposited in GenBank (www.<br />

ncbi.nlm.nih.gov); their accession numbers are listed in Table 1.<br />

Alignments <strong>and</strong> trees were deposited in TreeBASE (www.treebase.<br />

org).<br />

Table 1. (Page 158–160).<br />

a<br />

Abbreviations used: ATCC = American Type Culture Collection, Virginia, U.S.A.; <strong>CBS</strong> = Centraalbureau voor Schimmelcultures, Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; CPC = Culture<br />

Collection of Pedro Crous, housed at <strong>CBS</strong>, Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; dH = de Hoog Culture Collection, housed at <strong>CBS</strong>, Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; EXF = Culture Collection<br />

of Extremophilic Fungi, Ljubljana, Slovenia; IFO = Institute for Fermentation, Culture Collection of Microorganizms, Osaka, Japan; IMI = <strong>The</strong> International Mycological<br />

Institute, Egham, Surrey, U.K.; MZKI = Microbiological Culture Collection of the National Institute of Chemistry, Ljubljana, Slovenia; UAMH = University of Alberta Microfungus<br />

Collection, Alberta, Canada; VKM = All-Russian Collection of Microorganisms, Russian Academy of Sciences, Institute of Biochemistry <strong>and</strong> Physiology of Microorganisms,<br />

Pushchino, Russia; NT = ex-neotype strain; T = ex-type strain.<br />

b<br />

Reference: 1 de Hoog et al. 1999; 2 Park et al. 2004; 3 Braun et al. 2003; 4 Buzina et al. 2003; 5 Meklin et al. 2004; 6 Haubold et al. 1998; 7 Sert & Sterflinger, unpubl.; 8 Curtis et<br />

al. 1994; 9 Menkis et al. 2005; 10 Managbanag et al. unpubl.; 11 Crous et al. 2004; 12 Wirsel et al. 2002. All others are newly reported here.<br />

www.studiesinmycology.org<br />

161


Zalar et al.<br />

AY251097 <strong>Cladosporium</strong> uredinicola<br />

<strong>Cladosporium</strong> oxysporum <strong>CBS</strong> 125.80<br />

<strong>Cladosporium</strong> herbarum ATCC 66670<br />

<strong>Cladosporium</strong> bruhnei <strong>CBS</strong> 177.71<br />

<strong>Cladosporium</strong> spinulosum EXF-382<br />

<strong>Cladosporium</strong> psychrotolerans EXF-326<br />

<strong>Cladosporium</strong> psychrotolerans EXF-332<br />

<strong>Cladosporium</strong> langeronii <strong>CBS</strong> 189.54 T<br />

<strong>Cladosporium</strong> tenuissimum EXF-324<br />

<strong>Cladosporium</strong> cladosporioides <strong>CBS</strong> 170.54 NT<br />

AY251092 <strong>Cladosporium</strong> colocasiae<br />

AY251096 <strong>Cladosporium</strong> herbarum<br />

98<br />

<strong>Cladosporium</strong> halotolerans <strong>CBS</strong> 119416 T<br />

<strong>Cladosporium</strong> halotolerans EXF-228<br />

<strong>Cladosporium</strong> dominicanum <strong>CBS</strong> 119415 T<br />

<strong>Cladosporium</strong> velox <strong>CBS</strong> 119417 T<br />

66<br />

<strong>Cladosporium</strong> sphaerospermum EXF-464<br />

<strong>Cladosporium</strong> fusiforme <strong>CBS</strong> 119414 T<br />

<strong>Cladosporium</strong> sphaerospermum <strong>CBS</strong> 193.54 NT<br />

<strong>Cladosporium</strong> salinae <strong>CBS</strong> 119413 T<br />

AY251107 Pseudocercospora protearum<br />

80 AY251106 Pseudocercospora angolensis<br />

64<br />

AY251105 Cercospora cruenta<br />

AY251104 Cercospora zebrina<br />

AY251109 Passalora fulva<br />

79<br />

AY251108 Passalora dodonaeae<br />

AY251117 Mycosphaerella graminicola<br />

AY251110 Ramulispora sorghi<br />

100<br />

AY251114 Mycosphaerella latebrosa<br />

AY251102 Batcheloromyces proteae<br />

AY251101 Mycosphaerella lateralis<br />

76 AY251120 Teratosphaeria nubilosa<br />

AY251118 Catenulostroma macowanii<br />

AY251121 Devriesia staurophorum<br />

97 U42474 Dothidea insculpta<br />

93 AY016343 Dothidea ribesia<br />

AY016342 Discosphaerina fagi<br />

AY251126 Fusicladium effusum<br />

AY251122 Fusicladium amoenum Venturiaceae<br />

Dothioraceae & Dothideaceae<br />

10 steps<br />

Fig. 1. One of 30 equally most parsimonious <strong>and</strong> equally looking phylogenetic trees based on a heuristic tree search using aligned small subunit ribosomal DNA sequences.<br />

<strong>The</strong> tree was r<strong>and</strong>omly selected. Support based on 10 000 replicates of a fast step-wise addition bootstrap analysis is indicated near the branches. Species of <strong>Cladosporium</strong> s.<br />

str., including the seven newly described species, form a strongly supported monophyletic group among other taxa of the Mycosphaerellaceae (Dothideomycetes) (CI = 0.631,<br />

RI = 0.895, PIC = 50).<br />

Teratosphaeriaceae<br />

Davidiellaceae<br />

Mycosphaerellaceae<br />

Table 2. Statistical parameters describing phylogenetic analyses performed on sequence alignments of four different loci.<br />

Parameter SSU rDNA ITS rDNA 1 β-tubulin 2 Actin 3<br />

Number of alignment positions 1031 498 654 210<br />

Number of parsimony informative characters (PIC) 50 68 220 103<br />

Length of tree / number of steps 103 102 714 338<br />

Consistency Index (CI) 0.631 0.804 0.538 0.586<br />

Retention Index (RI) 0.895 0.975 0.883 0.885<br />

Rescaled Consistency Index (RC) 0.565 0.784 0.475 0.518<br />

Homoplasy index (HI) 0.369 0.196 0.462 0.414<br />

Number of equally parsimonious trees retained 30 600 90 32<br />

1<br />

Including the internal transcribed spacer region 1 <strong>and</strong> 2 <strong>and</strong> the 5.8S rDNA.<br />

2<br />

Including partial sequences of 4 exons <strong>and</strong> complete sequences of 3 introns.<br />

3<br />

Including partial sequences of 3 exons <strong>and</strong> 2 introns.<br />

162


<strong>Cladosporium</strong> sphaerospermum species complex<br />

69<br />

EXF-322<br />

<strong>CBS</strong> 119413<br />

EXF-604<br />

10 steps<br />

100<br />

85<br />

C. salinae<br />

62<br />

64<br />

79<br />

98<br />

EXF-780<br />

EXF-321<br />

EXF-946<br />

<strong>CBS</strong> 170.54<br />

ATCC 76499<br />

ATCC 66669<br />

EXF-699<br />

EXF-710<br />

EXF-697<br />

EXF-711<br />

<strong>CBS</strong> 125.80<br />

EXF-563<br />

EXF-452<br />

EXF-371<br />

Moricca et al.<br />

ATCC 38027<br />

EXF-714<br />

EXF-332<br />

EXF-391<br />

EXF-326<br />

87 <strong>CBS</strong> 109868<br />

dH 13833<br />

<strong>CBS</strong> 601.84<br />

dH 11736<br />

<strong>CBS</strong> 101880<br />

<strong>CBS</strong> 189.54<br />

dH 12459<br />

100 EXF-449<br />

<strong>CBS</strong> 452.71<br />

EXF-397<br />

EXF-466<br />

<strong>CBS</strong> 119417<br />

EXF-471<br />

EXF-739<br />

<strong>CBS</strong> 102045<br />

<strong>CBS</strong> 109.14<br />

EXF-455<br />

EXF-738<br />

EXF-962<br />

EXF-1061<br />

<strong>CBS</strong> 193.54<br />

EXF-458<br />

<strong>CBS</strong> 300.96 <strong>Cladosporium</strong> sp.<br />

EXF-732<br />

80<br />

90<br />

62<br />

100<br />

CPC 11683<br />

EXF-718<br />

EXF-720<br />

EXF-727<br />

72<br />

63<br />

79<br />

99<br />

EXF-977<br />

EXF-1072<br />

dH 12991<br />

dH 12862<br />

dH 13911<br />

EXF-564 C. halotolerans<br />

EXF-380<br />

EXF-703<br />

EXF-646<br />

EXF-228<br />

EXF-572<br />

<strong>CBS</strong> 280.49 “Mycosphaerella” hyperici<br />

<strong>CBS</strong> 177.71 C. bruhnei<br />

<strong>CBS</strong> 812.71<br />

EXF-595 <strong>Cladosporium</strong> sp.<br />

EXF-454 C. ramotenellum<br />

ATCC 66670 Davidiella tassiana<br />

EXF-333<br />

EXF-382<br />

EXF-334<br />

EXF-343 C. subinflatum<br />

C. spinulosum<br />

C. cladosporioides<br />

C. oxysporum<br />

C. tenuissimum<br />

C. sphaerospermum<br />

C. dominicanum<br />

C. psychrotolerans<br />

C. langeronii<br />

C. fusiforme<br />

C. velox<br />

Fig. 2. One of 600 equally most parsimonious <strong>and</strong> equally looking phylogenetic trees based on a heuristic tree search using aligned sequences of the internal transcribed<br />

spacer regions 1 <strong>and</strong> 2 <strong>and</strong> the 5.8S rDNA. <strong>The</strong> tree was r<strong>and</strong>omly selected. Support based on 10 000 replicates of a fast step-wise addition bootstrap analysis is indicated near<br />

the branches. Trees were rooted with the strains of <strong>Cladosporium</strong> salinae. Most monophyletic species clades received high, but some deeper branches moderate, bootstrap<br />

support (CI = 0.804, RI = 0.975, PIC = 68).<br />

RESULTS<br />

Descriptive statistical parameters of phylogenetic analyses <strong>and</strong><br />

calculated tree scores for each analysed sequence locus are<br />

summarised in Table 2. Mainly reference material such as extype<br />

or ex-neotype strains was analysed on the level of SSU<br />

rDNA sequences. Downloaded <strong>and</strong> newly generated SSU rDNA<br />

sequences of members of <strong>Cladosporium</strong> s. str. were compared<br />

with related taxa of the Mycosphaerellaceae, Dothioraceae <strong>and</strong><br />

Dothideaceae. <strong>The</strong> somewhat more distantly related Fusicladium<br />

effusum (Venturiaceae) (Braun et al. 2003: Fig. 2) was selected<br />

as outgroup. Anungitopsis amoena R.F. Castañeda & Dugan (now<br />

placed in Fusicladium Bonord., see Crous et al. 2007b), also a<br />

member of the Venturiaceae, was included in the analyses. All taxa<br />

included in the SSU rDNA analysis belong to the Dothideomycetes<br />

www.studiesinmycology.org<br />

(Schoch et al. 2006), within which the ingroup is represented by<br />

the orders Capnodiales (Davidiellaceae, Mycosphaerellaceae,<br />

Teratosphaeriaceae) <strong>and</strong> Dothideales (Dothioraceae, Dothideaceae)<br />

(see also Schoch et al. 2006). <strong>The</strong> <strong>genus</strong> <strong>Cladosporium</strong>, of which<br />

some species are linked to Davidiella Crous & U. Braun teleomorphs<br />

(Braun et al. 2003), forms a statistically strongly supported<br />

monophyletic group (Davidiellaceae). It also accommodates species<br />

newly described in this paper, namely, C. halotolerans Zalar, de<br />

Hoog & Gunde-Cimerman, C. fusiforme Zalar, de Hoog & Gunde-<br />

Cimerman, C. dominicanum Zalar, de Hoog & Gunde-Cimerman, C.<br />

salinae, C. psychrotolerans Zalar, de Hoog & Gunde-Cimerman, C.<br />

velox Zalar, de Hoog & Gunde-Cimerman <strong>and</strong> C. spinulosum Zalar,<br />

de Hoog & Gunde-Cimerman (Fig. 1). A sister group relationship<br />

of <strong>Cladosporium</strong> s. str. with a clade of taxa characterised,<br />

among others, by Mycosphaerella Johanson teleomorphs,<br />

containing various anamorphic genera such as Septoria Sacc.,<br />

163


Zalar et al.<br />

97 EXF-322<br />

<strong>CBS</strong> 119413<br />

EXF-604<br />

10 steps<br />

89<br />

C. salinae<br />

96<br />

61<br />

87<br />

93<br />

91<br />

98<br />

98 <strong>CBS</strong> 601.84<br />

<strong>CBS</strong> 101880<br />

dH 12459<br />

<strong>CBS</strong> 189.54<br />

<strong>CBS</strong> 109868<br />

100<br />

dH 13833<br />

dH 11736<br />

EXF-332<br />

EXF-326<br />

EXF-391<br />

EXF-714<br />

100 <strong>CBS</strong> 177.71 C. bruhnei<br />

ATCC 66670 Davidiella tassiana<br />

<strong>CBS</strong> 170.54 C. cladosporioides<br />

EXF-343 C. subinflatum<br />

EXF-382<br />

C. spinulosum<br />

EXF-334<br />

ATCC 66669<br />

<strong>CBS</strong> 125.80<br />

<strong>CBS</strong> 119417 C. velox<br />

C. oxysporum<br />

EXF-572<br />

EXF-564<br />

EXF-1072<br />

EXF-380<br />

C. halotolerans<br />

EXF-703<br />

EXF-228<br />

EXF-646<br />

EXF-977<br />

<strong>CBS</strong> 280.49 “Mycosphaerella” hyperici<br />

98<br />

dH 12862<br />

100<br />

63<br />

84<br />

100<br />

95 93 100<br />

100<br />

dH 12941<br />

dH 13911<br />

dH 12991<br />

EXF-727<br />

90 8862<br />

100<br />

99<br />

EXF-458<br />

98<br />

EXF-455<br />

<strong>CBS</strong> 102045<br />

EXF-1061<br />

<strong>CBS</strong> 193.54<br />

EXF-738<br />

<strong>CBS</strong> 109.14<br />

EXF-739<br />

EXF-962<br />

EXF-397<br />

EXF-449<br />

<strong>CBS</strong> 452.71<br />

78<br />

99<br />

EXF-720<br />

EXF-732<br />

EXF-718<br />

CPC 11683<br />

EXF-696<br />

C. halotolerans<br />

C. sphaerospermum<br />

C. fusiforme<br />

C. langeronii<br />

C. psychrotolerans<br />

C. dominicanum<br />

Fig. 3. One of 90 equally most parsimonious <strong>and</strong> equally looking phylogenetic trees based on a heuristic tree search using aligned exons <strong>and</strong> introns of a part of the β-tubulin<br />

gene. <strong>The</strong> tree was r<strong>and</strong>omly selected. Support based on 10 000 replicates of a fast step-wise addition bootstrap analysis is indicated near the branches. Trees were rooted with<br />

the strains of <strong>Cladosporium</strong> salinae. Most monophyletic species clades received high, but deeper branches weak or no, bootstrap support (CI = 0.538, RI = 0.883, PIC = 220).<br />

Ramularia Unger, Cercospora Fresen., Pseudocercospora Speg.,<br />

“Trimmatostroma” Corda (now Catenulostroma Crous & U. Braun)<br />

(see Crous et al. 2004, 2007a – this volume) <strong>and</strong> the somewhat<br />

cladosporium-like <strong>genus</strong> Devriesia Seifert & N.L. Nick. (Seifert et al.<br />

2004), was statistically only moderately supported (Fig. 1), whereas<br />

in an analogous analysis by Braun et al. (2003: Fig. 2) it was highly<br />

supported. <strong>The</strong>se data also support the conclusion by Braun et al.<br />

(2003) <strong>and</strong> Crous et al. (2006) that <strong>Cladosporium</strong> is not a member<br />

of the distantly related Herpotrichiellaceae (Chaetothyriomycetes),<br />

which is also rich in cladosporium-like taxa (Crous et al. 2006).<br />

None of the fungi isolated from hypersaline environments belonged<br />

to the Herpotrichiellaceae. <strong>The</strong> SSU rDNA sequences do not<br />

resolve a phylogenetic structure within <strong>Cladosporium</strong> s. str. Only<br />

a moderately supported clade comprising C. halotolerans, C.<br />

dominicanum, C. velox, C. sphaerospermum <strong>and</strong> C. fusiforme is<br />

somewhat distinguished from a statistically unsupported clade with<br />

C. herbarum (Pers. : Fr.) Link, C. cladosporioides (Fresen.) G.A.<br />

de Vries, C. oxysporum Berk. & Broome, C. spinulosum, <strong>and</strong> C.<br />

psychrotolerans, etc. Because C. salinae appeared most distinct<br />

within the <strong>genus</strong> <strong>Cladosporium</strong> in analyses of the SSU rDNA (Fig.<br />

1), it was used as outgroup in analyses of the ITS rDNA <strong>and</strong> the<br />

β-tubulin <strong>and</strong> actin genes.<br />

Analyses of the more variable ITS rDNA <strong>and</strong> partial β-tubulin<br />

<strong>and</strong> actin gene introns <strong>and</strong> exons supported the species clades of<br />

C. halotolerans, C. dominicanum, C. sphaerospermum, C. fusiforme<br />

<strong>and</strong> C. velox (Figs 2–4), of which C. velox was distinguished in<br />

the β-tubulin tree by a particular long terminal branch of the only<br />

sequenced strain (Fig. 3). <strong>Cladosporium</strong> salinae also clustered as a<br />

well-supported species clade in preliminary analyses using various<br />

Mycosphaerella species as outgroup (not shown). All strains of C.<br />

langeronii (Fonseca, Leão & Nogueira) Vuill. are particularly well<br />

distinguishable from other <strong>Cladosporium</strong> species by strikingly slow-<br />

164


<strong>Cladosporium</strong> sphaerospermum species complex<br />

EXF-322<br />

<strong>CBS</strong> 119413<br />

EXF-604<br />

60<br />

100<br />

C. salinae<br />

<strong>CBS</strong> 189.54<br />

71 <strong>CBS</strong> 601.84<br />

<strong>CBS</strong> 101880<br />

dH 12459<br />

dH 13833<br />

96 <strong>CBS</strong> 109868<br />

dH 11736<br />

62 EXF-332<br />

EXF-391<br />

77 EXF-714<br />

<strong>CBS</strong> 177.71 C. bruhnei<br />

66 89 ATCC 66670 Davidiella tassiana<br />

67 EXF-343 C. subinflatum<br />

85<br />

C. psychrotolerans<br />

87 EXF-696<br />

EXF-718<br />

100<br />

EXF-732<br />

C. dominicanum<br />

CPC 11683<br />

EXF-598<br />

100<br />

<strong>CBS</strong> 102045<br />

EXF-458<br />

EXF-1069<br />

EXF-455<br />

72<br />

EXF-1061<br />

77EXF-739<br />

76<br />

EXF-962<br />

C. sphaerospermum<br />

81 EXF-738<br />

76 <strong>CBS</strong> 109.14<br />

<strong>CBS</strong> 193.54<br />

<strong>CBS</strong> 300.96 <strong>Cladosporium</strong> sp.<br />

EXF-449<br />

100 90<br />

EXF-397 C. fusiforme<br />

<strong>CBS</strong> 452.71<br />

100<br />

EXF-466<br />

EXF-471<br />

<strong>CBS</strong> 119417<br />

C. velox<br />

84 98 EXF-334<br />

EXF-382<br />

C. spinulosum<br />

<strong>CBS</strong> 170.54 C. cladosporioides<br />

89<br />

<strong>CBS</strong> 125.80<br />

ATCC 66669<br />

C. oxysporum<br />

dH 12862<br />

dH 13911<br />

EXF-646<br />

EXF-703<br />

EXF-228<br />

EXF-1072 C. halotolerans<br />

EXF-572<br />

EXF-977<br />

EXF-564<br />

EXF-380<br />

<strong>CBS</strong> 280.49 “Mycosphaerella” hyperici<br />

C. langeronii<br />

10 steps<br />

Fig. 4. One of 32 equally most parsimonious <strong>and</strong> equally looking phylogenetic trees based on a heuristic tree search using aligned exons <strong>and</strong> introns of the partial actin gene.<br />

<strong>The</strong> tree was r<strong>and</strong>omly selected. Support based on 10 000 replicates of a fast step-wise addition bootstrap analysis is indicated near the branches. Trees were rooted with<br />

the strains of <strong>Cladosporium</strong> salinae. Most monophyletic species clades received high, but deeper branches weak or no, bootstrap support (CI = 0.586, RI = 0.885, PIC = 103).<br />

growing colonies at all tested temperatures <strong>and</strong> relatively large,<br />

oblong conidia. However, phylogenetic analyses of the β-tubulin <strong>and</strong><br />

actin gene indicate that C. langeronii presents two cryptic species<br />

(Figs 3–4). <strong>The</strong> species clade of C. psychrotolerans is moderately<br />

supported in analyses of the actin gene but highly by means of<br />

the β-tubulin gene. It is evident from all three analyses (Figs 2–<br />

4) that C. langeronii <strong>and</strong> C. psychrotolerans are closely related<br />

species. <strong>The</strong> species node of <strong>Cladosporium</strong> spinulosum, which is<br />

morphologically clearly distinguished from all other species by its<br />

conspicuous ornamentation consisting of digitate projections (Fig.<br />

5), is supported by β-tubulin (Fig. 3) <strong>and</strong> actin (Fig. 4) sequence<br />

data but not by those of the ITS rDNA (Fig. 2). Analyses of all loci,<br />

however, indicate that it is a member of the C. herbarum complex.<br />

<strong>The</strong> analyses of sequences of the ITS <strong>and</strong> the β-tubulin<br />

<strong>and</strong> actin gene introns <strong>and</strong> exons (Figs 2–4) do not allow the<br />

full elucidation of phylogenetic relationships among these<br />

<strong>Cladosporium</strong> species. Statistical support of the interior tree<br />

branches resulting from analyses of the β-tubulin <strong>and</strong> actin genes<br />

is low (bootstrap values mostly < 50 %). While the sister group<br />

relationship of C. sphaerospermum <strong>and</strong> C. fusiforme is highly<br />

supported in the analysis based on the β-tubulin gene, analysis<br />

of the ITS rDNA indicate that these two species are unrelated, <strong>and</strong><br />

that C. sphaerospermum is closely related to C. dominicanum. It is<br />

clear from the data that the species morphologically resembling C.<br />

sphaerospermum are not phylogenetically closely related <strong>and</strong> that<br />

the data we present here do not allow their classification in natural<br />

subgroups of the <strong>genus</strong> <strong>Cladosporium</strong>. Only C. spinulosum was<br />

placed in all analyses among species of the C. herbarum complex<br />

www.studiesinmycology.org<br />

165


Zalar et al.<br />

Fig. 5. Conidial scars <strong>and</strong> surface ornamentation of ramoconidia <strong>and</strong> conidia (SEM). A. C. dominicanum (strain EXF-732). B. C. fusiforme (strain EXF-449). C. C. halotolerans<br />

(strain EXF-572). D. C. langeronii (strain <strong>CBS</strong> 189.54). E. C. psychrotolerans (strain EXF-391). F. C. salinae (strain EXF-335 = <strong>CBS</strong> 119413). G. C. sphaerospermum (strain<br />

<strong>CBS</strong> 193.54). H. C. spinulosum (strain EXF-334). I. C. velox (strain <strong>CBS</strong> 119417). Scale bars = 5 µm. (Photos: K. Drašlar).<br />

<strong>and</strong> all analyses supported close relatedness of C. langeronii <strong>and</strong><br />

C. psychrotolerans.<br />

<strong>The</strong> majority of species described here have slightly<br />

ornamented conidia ranging from minutely verruculose (C.<br />

fusiforme, C. langeronii, C. psychrotolerans, C. sphaerospermum,<br />

C. velox) to verrucose (C. halotolerans) (Fig. 5). <strong>The</strong> verrucose<br />

conidia of C. halotolerans can be recognised also under the<br />

light microscope <strong>and</strong> used as a distinguishing character. Almost<br />

smooth to minutely verruculose conidia are encountered in C.<br />

dominicanum <strong>and</strong> C. salinae (Fig. 5). <strong>Cladosporium</strong> spinulosum,<br />

a member of the C. herbarum species complex, has conidia with<br />

a digitate ornamentation that can appear spinulose under the light<br />

microscope; however, when using the SEM it became clear that its<br />

projections have parallel sides <strong>and</strong> a blunt end (Fig. 5).<br />

DISCUSSION<br />

<strong>The</strong> <strong>genus</strong> <strong>Cladosporium</strong> was established by Link (1816) who<br />

originally included four species, of which C. herbarum is the<br />

type species of the <strong>genus</strong> (Clements & Shear 1931). In 1950,<br />

von Arx reported a teleomorph connection for this species with<br />

Mycosphaerella tassiana (De Not.) Johanson. Based on SSU<br />

rDNA data the majority of Mycosphaerella species, including<br />

the type species of the <strong>genus</strong>, M. punctiformis (Pers.) Starbäck,<br />

clustered within the Mycosphaerellaceae, a family separated<br />

from M. tassiana (Braun et al. 2003). <strong>The</strong>refore, Mycosphaerella<br />

tassiana was reclassified as Davidiella tassiana (De Not.) Crous<br />

& U. Braun, the type of the new <strong>genus</strong> Davidiella. All anamorphs<br />

with a cladosporium- <strong>and</strong> heterosporium-like appearance <strong>and</strong> with<br />

a supposed Dothideomycetes relationship were maintained under<br />

the anamorph name <strong>Cladosporium</strong>, morphologically characterised<br />

by scars with a protuberant hilum consisting of a central dome<br />

surrounded by a raised rim (David 1997).<br />

<strong>The</strong> concept of distinguishing ramoconidia from secondary<br />

ramoconidia has been adopted from Schubert et al. (2007). In<br />

the species described here, ramoconidia have been observed<br />

often in C. sphaerospermum, sometimes in C. psychrotolerans,<br />

C. langeronii <strong>and</strong> C. spinulosum, <strong>and</strong> only sporadically in all other<br />

species. <strong>The</strong>refore, ramoconidia can be seen as important for<br />

distinguishing species although sometimes, they can be observed<br />

only with difficulty. When using ramoconidia as a diagnostic<br />

criterion, colonies only from SNA <strong>and</strong> not older than 7 d should be<br />

taken into account.<br />

<strong>Cladosporium</strong> sphaerospermum was described by Penzig<br />

(1882) from decaying Citrus leaves <strong>and</strong> branches in Italy. He<br />

described C. sphaerospermum as a species with (i) branched,<br />

septate <strong>and</strong> dark conidiophores having a length of 150–300 µm <strong>and</strong><br />

a width of the main conidiophore stipe of 3.5–4 µm, (ii) spherical to<br />

ellipsoid, acrogenously formed conidia of 3.4–4 µm diam, <strong>and</strong> (iii)<br />

ramoconidia of 6–14 × 3.5–4 µm. Penzig’s original material is not<br />

known to be preserved. Later, a culture derived from <strong>CBS</strong> 193.54,<br />

originating from a human nail, was accepted as typical of C.<br />

sphaerospermum. However, de Vries (1952), incorrectly cited it as<br />

“lectotype”, <strong>and</strong> thus the same specimen is designated as neotype<br />

in this study (see below), with the derived culture (<strong>CBS</strong> 193.54)<br />

166


<strong>Cladosporium</strong> sphaerospermum species complex<br />

used as ex-neotype strain. Numerous strains with identical or very<br />

<strong>similar</strong> ITS rDNA sequences as <strong>CBS</strong> 193.54 were isolated from<br />

hypersaline water or organic substrata including plants or walls of<br />

bathrooms. It is not clear yet whether surfaces in bathrooms <strong>and</strong> of<br />

plants, colonised by C. sphaerospermum, can have a <strong>similar</strong> low<br />

water activity as salterns. In our experiments, the strains of this<br />

species, however, grew under in vitro conditions at a water activity<br />

of up to 0.860, while Hocking et al. (1994) <strong>and</strong> Aihara et al. (2002)<br />

reported that it can grow even at 0.815. <strong>The</strong>refore, we consider<br />

C. sphaerospermum as halo- or osmotolerant. Hardly any reports<br />

are available unambiguously proving that C. sphaerospermum is<br />

a human pathogen. It is therefore possible that <strong>CBS</strong> 193.54 was<br />

not involved in any disease process but rather occurred as a<br />

contaminant on dry nail material. <strong>Cladosporium</strong> sphaerospermum<br />

is a phylogenetically well-delineated species (Figs 2–4).<br />

Strains of C. halotolerans were isolated sporadically from<br />

substrata such as peanut cell suspension, tissue culture, bathroom<br />

walls <strong>and</strong> as culture contaminants. This surprising heterogeneity<br />

of substrata suggests that C. halotolerans is distributed by air <strong>and</strong><br />

that it can colonise whatever substrata available, although it may<br />

have its natural niche elsewhere. We have recurrently isolated it<br />

from hypersaline water of salterns <strong>and</strong> other saline environments<br />

<strong>and</strong> it was also detected with molecular methods (but not isolated)<br />

from skin of a salt water dolphin. <strong>The</strong>re are only few reports of<br />

this species from plants (Table 1). It is therefore possible that C.<br />

halotolerans is a species closely linked to salty or hypersaline<br />

environments although additional sampling is necessary to prove<br />

that. <strong>Cladosporium</strong> halotolerans is morphologically recognisable<br />

by relatively oblong to spherical, coarsely rough-walled conidia.<br />

<strong>The</strong> ITS rDNA sequence of a fungus in the skin of a bottlenose<br />

dolphin, suffering from lobomycosis, is identical to the sequences<br />

of C. halotolerans. This sequence was deposited as Lacazia<br />

loboi Taborda, V.A. Taborda & McGinnis (GenBank AF035674) by<br />

Haubold et al. (1998), who apparently concluded wrongly that a<br />

fungus with a cladosporium-like ITS rDNA sequence <strong>similar</strong> to that<br />

of C. halotolerans can be the agent of lobomycosis. Later, Herr<br />

et al. (2001) showed that Lacazia loboi phylogenetically belongs<br />

to the Onygenales on the basis of amplified SSU rDNA <strong>and</strong> chitin<br />

synthase-2 gene sequences generated from tissue lesions. By<br />

this, they confirmed an earlier supposition by Lacaz (1996) who<br />

reclassified the organism as Paracoccidioides loboi O.M. Fonseca<br />

& Silva Lacaz (Onygenales). It is therefore possible that C.<br />

halotolerans was not the main etiologic agent for the lobomycosis<br />

<strong>and</strong> it was colonising the affected dolphin skin secondarily while<br />

inhabiting other seawater habitats.<br />

<strong>Cladosporium</strong> langeronii <strong>and</strong> C. psychrotolerans are closely<br />

related but C. langeronii is particularly well distinguishable from<br />

all other <strong>Cladosporium</strong> species by its slow growing colonies<br />

(1–7 mm diam / 14 d) <strong>and</strong> relatively large conidia (4–5.5 × 3–4<br />

μm). <strong>Cladosporium</strong> psychrotolerans has smaller conidia (3–4 ×<br />

2.5–3 μm) but a <strong>similar</strong> length : width ratio <strong>and</strong> faster exp<strong>and</strong>ing<br />

colonies (8–18 mm diam / 14 d). <strong>Cladosporium</strong> langeronii is most<br />

likely a complex of at least two species. Strains isolated from the<br />

Arctic <strong>and</strong> the Antarctic may need to be distinguished from C.<br />

langeronii s. str. on species level. This inference is particularly<br />

supported by analyses of the β-tubulin <strong>and</strong> actin genes (Figs 3–4).<br />

<strong>Cladosporium</strong> langeronii s. str., represented by an authentic strain<br />

of Hormodendrum langeronii Fonseca, Leão & Nogueira, <strong>CBS</strong><br />

189.54 (Trejos 1954), has been isolated from a variety of substrata<br />

but is tolerating only up to 10 % NaCl. It was originally described<br />

by da Fonseca et al. (1927a, b) <strong>and</strong> subsequently reclassified<br />

as <strong>Cladosporium</strong> langeronii by Vuillemin (1931). <strong>The</strong> authentic<br />

www.studiesinmycology.org<br />

strain derived from an ulcerating nodular lesion on the arm of a<br />

human patient. Because other strains of this species are ubiquitous<br />

saprobes originating from various substrata, we suspect that C.<br />

langeronii is not an important human pathogen. <strong>Cladosporium</strong><br />

psychrotolerans has been isolated from hypersaline environments<br />

only, <strong>and</strong> tolerates up to 20 % NaCl in culture media.<br />

In general, the human- or animal-pathogenic role of the C.<br />

sphaerospermum-like species described here seems to be limited.<br />

It is possible that pathogenic species of Cladophialophora Sacc.<br />

have been misidentified as C. sphaerospermum or as other<br />

species of <strong>Cladosporium</strong> (de Hoog et al. 2000). Alternatively, true<br />

<strong>Cladosporium</strong> species isolated as clinical strains could have been<br />

secondary colonisers since they are able to dwell on surfaces poor<br />

in nutrients, possibly in an inconspicuous dormant phase <strong>and</strong> may<br />

then be practically invisible. More likely, they could be air-borne<br />

contaminations of lesions, affected nails etc. (Summerbell et al.<br />

2005) or are perhaps disseminated by insufficiently sterilised medical<br />

devices, as melanised fungi can be quite resistant to disinfectants<br />

(Phillips et al. 1992). <strong>The</strong>y can easily be isolated <strong>and</strong> rapidly<br />

become preponderant at isolation <strong>and</strong> thus difficult to exclude as<br />

etiologic agents of a disease. For example, in 2002, a case report<br />

on an intrabronchial lesion by C. sphaerospermum in a healthy,<br />

non-asthmatic woman was described (Yano et al. 2002), but we<br />

judge the identification of the causal agent to remain uncertain, as<br />

it was based on morphology alone <strong>and</strong> no culture is available. <strong>The</strong><br />

present authors have the opinion that all clinical cases ascribed to<br />

<strong>Cladosporium</strong> species need careful re-examination.<br />

General characteristics <strong>and</strong> description of <strong>Cladosporium</strong><br />

sphaerospermum-like species<br />

<strong>The</strong> present paper focuses on <strong>Cladosporium</strong> strains isolated from<br />

hypersaline environments. Comparison of data from deliberate<br />

sampling <strong>and</strong> analysis of reference strains from culture collections<br />

inevitably leads to statistical bias, <strong>and</strong> therefore a balanced<br />

interpretation of ecological preferences of the species presented is<br />

impossible. Nevertheless, some species appeared to be consistent<br />

in their choice of habitat, <strong>and</strong> for this reason we summarise<br />

isolation data for all species described. Strains belonging to a single<br />

molecular clade proved to have <strong>similar</strong> cultural characteristics <strong>and</strong><br />

microscopic morphology. Although within most of the species there<br />

was some molecular variation noted (particularly when intron-rich<br />

genes were analysed), some consistent phenetic trends could be<br />

observed.<br />

Conidiophores of all C. sphaerospermum-like species lack<br />

nodose inflations (McKemy & Morgan-Jones 1991). <strong>The</strong>y are<br />

usually ascending <strong>and</strong> can sometimes be poorly differentiated from<br />

their supporting hyphae. Though the initiation point of conidiophore<br />

stipes could sometimes be determined only approximately,<br />

their lengths were in some cases useful for distinguishing<br />

morphologically <strong>similar</strong> species when observed in slide cultures.<br />

Generally, the branched part of a conidiophore forms a complex<br />

tree-like structure. <strong>The</strong> number <strong>and</strong> orientation of early formed<br />

secondary ramoconidia, however, determines whether it is<br />

rotationally symmetric or unilateral.<br />

<strong>The</strong> variability in ITS rDNA sequences observed in all C.<br />

sphaerospermum-like species (about 10 %) spans the variation<br />

observed in all members of the <strong>genus</strong> <strong>Cladosporium</strong> sequenced<br />

to date. Thus, the C. sphaerospermum-like species described here<br />

may not present a single monophyletic group but may belong to<br />

various species complexes within <strong>Cladosporium</strong>. Verifying existing<br />

literature with sequence data of these species (Wirsel et al. 2002,<br />

Park et al. 2004), we noticed that names of the common saprobes<br />

seem to be distributed nearly at r<strong>and</strong>om over phylogenetic trees.<br />

167


Zalar et al.<br />

For most commonly used names, no type material is available for<br />

sequencing. Also verification of published reports is difficult without<br />

available voucher strains.<br />

<strong>Cladosporium</strong> cladosporioides was incorrectly lectotypified<br />

based on <strong>CBS</strong> 170.54 (de Vries 1952), which Bisby considered a<br />

st<strong>and</strong>ard culture of C. herbarum. <strong>The</strong> C. cladosporioides species<br />

complex requires revision, <strong>and</strong> will form the basis of a future study.<br />

<strong>Cladosporium</strong> herbarum is maintained as a dried specimen in the<br />

Leiden herbarium; Prasil & de Hoog (1988) selected <strong>CBS</strong> 177.71<br />

as a representative living strain. Strains, earlier accepted as living<br />

representatives of C. herbarum, <strong>CBS</strong> 177.71 <strong>and</strong> <strong>CBS</strong> 812.71<br />

(Prasil & de Hoog 1988, Wirsel et al. 2002) <strong>and</strong> ATCC 66670<br />

(Braun et al. 2003, as Davidiella tassiana) have been re-identified<br />

as C. bruhnei Linder by Schubert et al. (2007 – this volume). Ho<br />

et al. (1999) used strain ATCC 38027 as a representative of C.<br />

tenuissimum Cooke <strong>and</strong> this strain has identical ITS sequences as<br />

the non-deposited C. tenuissimum material used by Moricca et al.<br />

(1999). We tentatively accept this concept although we could not<br />

include ATCC 38027 in our analyses. <strong>The</strong> ITS sequence of strain<br />

<strong>CBS</strong> 125.80, identified by Wirsel et al. (2002) as C. oxysporum,<br />

is identical to the sequence of ATCC 38027. Strain ATCC 76499,<br />

published by Ho et al. (1999) as C. oxysporum, appears to be<br />

identical to a number of currently unidentified <strong>Cladosporium</strong> strains<br />

from Slovenian salterns that compose a cluster separate from all<br />

remaining species. Strains of this cluster, represented in Fig. 2 by<br />

strain ATCC 76499, morphologically resemble C. oxysporum.<br />

Strain <strong>CBS</strong> 300.96 has not been identified to species level<br />

in the present study. It clusters outside the species clade of<br />

C. sphaerospermum, with the latter being its nearest relative.<br />

<strong>CBS</strong> 300.96 differs from C. sphaerospermum by having smaller<br />

structures: conidiophore stipes [(5–)20–80(–150) × (2–)2.5–3(–4)<br />

μm], 0–1 septate ramoconidia [(13–)19–27(–32) × 2–2.5 μm],<br />

conidia [(2.5–)3–3.5(–4) × (2–)2–2.5(–3) μm] <strong>and</strong> secondary<br />

ramoconidia [(5–)9–18(–30) × (2–)2.5–2.5(–3) μm]. However,<br />

based on a single isolate, we currently refrain from describing it as<br />

a new species.<br />

Key to species treated in this study<br />

Macro-morphological characters used in the key are from colonies grown on PDA <strong>and</strong> MEA 14 d at 25 °C, if not stated otherwise; microscopical<br />

characters are from SNA slide cultures grown for 7 d at 25 °C.<br />

1. Conidial ornamentation conspicuously echinulate / digitate because of up to 1.3 µm long projections that have more or less parallel sides<br />

............................................................................................................................................................................................. C. spinulosum<br />

1. Conidial ornamentation verruculose to verrucose or smooth, not conspicuously echinulate or digitate .................................................... 2<br />

2. Conidiophores micronematous, poorly differentiated, once or several times geniculate-sinuous, short, up to 60 µm long; terminal conidia<br />

obovoid ....................................................................................................................................................................................... C. salinae<br />

2. Conidiophores micro- or macronematous, not geniculate or only slightly so, usually up to 100 µm or 220 µm long or even longer; terminal<br />

conidia globose, subglobose to ovoid or fusiform ...................................................................................................................................... 3<br />

3. Secondary ramoconidia 0–3(–4)-septate; septa of conidiophores <strong>and</strong> conidia darkened <strong>and</strong> thickened .................................................. 4<br />

3. Secondary ramoconidia 0–1(–2)-septate; septa neither darkened nor thickened ..................................................................................... 5<br />

4. Conidiophores (5–)10–50(–300) × (2–)2.5–3(–5.5) μm; terminal conidia (2–)3–4(–6) × (2–)2.5–3(–5) μm; secondary ramoconidia (5–)7–<br />

12(–37.5) × (2–)2.5–3(–6.5) μm; ramoconidia sporadically formed ................................................................................... C. halotolerans<br />

4. Conidiophores mostly longer <strong>and</strong> somewhat wider, (10–)45–130(–300) × (2.5–)3–4(–6) μm; terminal conidia mostly wider, (2.5–)3–4(–7)<br />

× (2–)3–3.5(–4.5) μm; secondary ramoconidia (4–)8.5–16(–37.5) × (2–)3–3.5(–5) μm; ramoconidia often formed, up to 40 µm long, with<br />

up to 5 septa ............................................................................................................................................................. C. sphaerospermum<br />

5. Terminal conidia usually fusiform ............................................................................................................................................ C. fusiforme<br />

5. Terminal conidia globose, subglobose or ovoid ......................................................................................................................................... 6<br />

6. Conidia <strong>and</strong> secondary ramoconidia irregularly verruculose to sometimes loosely verrucose; radial growth on PDA at 25 °C after 14 d<br />

typically less than 5 mm ......................................................................................................................................................... C. langeronii<br />

6. Conidia <strong>and</strong> secondary ramoconidia smooth to minutely verruculose; radial growth on PDA at 25 °C after 14 d typically more than 10 mm<br />

.................................................................................................................................................................................................................... 7<br />

7. Conidiophores (3–)3.5–4(–7.5) μm wide, thick-walled; conidiogenous loci <strong>and</strong> conidial hila 0.5–2 μm diam; ramoconidia sometimes<br />

formed with a broadly truncate, up to 2 µm wide non-cladosporioid base; no growth observed after 14 d at 30 °C on MEA<br />

..................................................................................................................................................................................... C. psychrotolerans<br />

7. Conidiophores mostly narrower, 2–4 μm wide, only with slightly thickened walls; conidiogenous loci <strong>and</strong> conidial hila narrower, 0.5–1.5<br />

μm diam; ramoconidia rarely formed; colony showing at least weak growth after 14 d at 30 °C on MEA ................................................. 8<br />

8. Secondary ramoconidia (4–)6.5–13(–24.5) μm long; no visible colony growth after 14 d at 10 °C on MEA.................... C. dominicanum<br />

8. Secondary ramoconidia mostly longer, (3.5–)5.5–19(–42) μm; radial growth of colonies after 14 d at 10 °C on MEA more than 5 mm<br />

........................................................................................................................................................................................................ C. velox<br />

168


<strong>Cladosporium</strong> sphaerospermum species complex<br />

Description of <strong>Cladosporium</strong> species<br />

<strong>Cladosporium</strong> dominicanum Zalar, de Hoog & Gunde-Cimerman,<br />

sp. nov. MycoBank MB510995. Fig. 6.<br />

Etymology: Refers to the the Dominican Republic, where most<br />

strains were encountered.<br />

Conidiophora lateralia vel terminalia ex hyphis rectis oriunda; stipes longitudine<br />

variabili, (5–)10–100(–200) × (1.5–)2–2.5(–3.5) μm, olivaceo-brunneus, levis vel<br />

leniter verruculosus, tenuitunicatus, plerumque unicellularis, simplex vel ramosus.<br />

Conidiorum catenae undique divergentes, ad 8 conidia in parte continua continentes.<br />

Cellulae conidiogenae indistinctae. Conidia levia vel leniter verruculosa, dilute<br />

brunnea, unicellularia, plerumque breviter ovoidea, utrinque angustata, (2.5–)<br />

3–3.5(–5.5) × (2–)2–2.5(–2.5) μm, long.: lat. 1.4–1.6; ramoconidia secundaria<br />

cylindrica vel quasi globosa, 0–1-septata, (4–)6.5–13(–24.5) × (2–)2.5–3(–4.5) μm,<br />

ad 4 cicatrices terminales ferentia; cicatrices inspissatae, protuberantes, 0.5–1.2 μm<br />

diam. Hyphae vagina polysaccharidica carentes.<br />

Mycelium without extracellular polysaccharide-like material.<br />

Conidiophores arising laterally <strong>and</strong> terminally on erect hyphae,<br />

micronematous <strong>and</strong> semimacronematous, stipes of variable length,<br />

(5–)10–100(–200) × (1.5–)2–2.5(–3.5) μm, olivaceous-brown,<br />

smooth to minutely verruculose, thin-walled, almost non-septate,<br />

unbranched or branched. Conidial chains branching in all directions,<br />

up to eight conidia in the unbranched parts. Conidiogenous cells<br />

undifferentiated. Ramoconidia rarely formed. Conidia smooth<br />

to minutely verruculose, subhyaline to light brown, non-septate,<br />

usually short-ovoid, narrower at both ends, length : width ratio =<br />

1.4–1.6; (2.5–)3–3.5(–5.5) × (2–)2–2.5(–2.5) μm [av. (± SD) 3.4<br />

(± 0.6) × 2.2 (± 0.2)]; secondary ramoconidia cylindrical to almost<br />

spherical, 0–1-septate, (4–)6.5–13(–24.5) × (2–)2.5–3(–4.5) μm<br />

[av. (± SD) 10.3 (± 5.2) × 2.7 (± 0.6)], with up to four distal scars.<br />

Conidiogenous scars thickened <strong>and</strong> conspicuous, protuberant,<br />

0.5–1.2 μm diam.<br />

Cultural characteristics: Colonies on PDA reaching 18–36 mm<br />

diam, olive-yellow (2D6), hairy granular, flat or slightly furrowed,<br />

with flat margin. Droplets of light reseda-green (2E6) exudate<br />

sometimes present. Reverse dark green to black. Colonies on OA<br />

reaching 19–34 mm diam, olive (2F5), loosely powdery with raised<br />

central part due to fasciculate bundles of conidiophores. Reverse<br />

dark green. Colonies on MEA reaching 30–32 mm diam, reseda<br />

green (2E6), velvety, furrowed, with undulate margin. Reverse dark<br />

green-brown. Colonies on MEA + 5 % NaCl reaching 37–41 mm<br />

diam, reseda-green (2E6), radially furrowed, velvety, sporulating in<br />

the central part or all over the colony, margin white <strong>and</strong> regular.<br />

Reverse brownish green.<br />

Maximum tolerated salt concentration: 75 % of tested strains<br />

develop colonies at 20 % NaCl after 7 d, while after 14 d all strains<br />

grow <strong>and</strong> sporulate.<br />

Cardinal temperatures: No growth at 4 <strong>and</strong> 10 °C, optimum 25 °C<br />

(30–32 mm diam), maximum 30 °C (2–15 mm diam), no growth at<br />

37 °C.<br />

Specimen examined: Dominican Republic, from hypersaline water of salt lake<br />

Enriquillo, coll. Nina Gunde-Cimerman, Jan. 2001, isol. P. Zalar 25 Feb. 2001, <strong>CBS</strong><br />

H-19733, holotype, culture ex-type EXF-732 = <strong>CBS</strong> 119415.<br />

Habitats <strong>and</strong> distribution: Fruit surfaces; hypersaline waters in<br />

(sub)tropical climates.<br />

Differential parameters: No growth at 10 °C, ovoid conidia, large<br />

amounts of sterile mycelium.<br />

Strains examined: CPC 11683, EXF-696, EXF-718, EXF-720, EXF-<br />

727, EXF-732 (= <strong>CBS</strong> 119415; ex-type strain).<br />

www.studiesinmycology.org<br />

Note: Cultures of C. dominicanum sporulate less abundantly than<br />

C. sphaerospermum <strong>and</strong> C. halotolerans <strong>and</strong> tend to lose their<br />

ability to sporulate with subculturing.<br />

<strong>Cladosporium</strong> fusiforme Zalar, de Hoog & Gunde-Cimerman, sp.<br />

nov. MycoBank MB510997. Fig. 7.<br />

Etymology: Refers to its usually fusiform conidia.<br />

Conidiophora erecta, lateralia vel terminalia ex hyphis rectis oriunda; stipes<br />

longitudine variabili, (10–)25–50(–100) × (2–)2–3.5(–4) μm, olivaceo-brunneus,<br />

levis, crassitunicatus, compluries septatus (cellulis 9–23 μm longis), plerumque<br />

simplex. Conidiorum catenae undique divergentes, in parte continua ad 5 conidia<br />

continentes. Cellulae conidiogenae indistinctae. Conidia leniter verruculosus, dilute<br />

brunnea, unicellularia, plerumque fusiformia, utrinque angustata, (2.5–)3.5–5(–<br />

6.5) × (2–)2–2.5(–3) μm, long. : lat. 1.8–2.0; ramoconidia secundaria cylindrica,<br />

0(–1)-septata, (5–)6–11(–22) × (2.5–)2.5–3(–3) μm, ad 4 cicatrices terminales<br />

ferentia; cicatrices inspissatae, conspicuae, 0.7–1.0 μm diam. Hyphae vagina<br />

polysaccharidica carentes.<br />

Mycelium without extracellular polysaccharide-like material.<br />

Conidiophores erect, arising laterally <strong>and</strong> terminally from straight<br />

hyphae, stipes of variable length, (10–)25–50(–100) × (2–)2–3.5(–<br />

4) μm, olivaceous-brown, smooth- <strong>and</strong> thick-walled, regularlyseptate<br />

(cell length 9–23 μm), mostly unbranched. Conidial chains<br />

branching in all directions, up to 5 conidia in the unbranched parts.<br />

Conidiogenous cells undifferentiated. Ramoconidia rarely formed.<br />

Conidia minutely verruculose, light brown, aseptate, usually fusiform<br />

<strong>and</strong> narrower at both ends, length : width ratio = 1.8–2.0; (2.5–)3.5–<br />

5(–6.5) × (2–)2–2.5(–3) μm [av. (± SD) 4.4 (± 0.8) × 2.2 (± 0.2)];<br />

secondary ramoconidia cylindrical, 0(–1)-septate, (5–)6–11(–22) ×<br />

(2.5–)2.5–3(–3) μm [av. (± SD) 9.0 (± 4.7) × 2.6 (± 0.3)], with up<br />

to 4 distal scars. Conidiogenous scars thickened <strong>and</strong> conspicuous,<br />

protuberant, 0.7–1.0 μm diam.<br />

Cultural characteristics: Colonies on PDA reaching 20–26 mm diam,<br />

dull green (30E3), granular due to profuse sporulation, flat, with flat<br />

margin. Sterile mycelium absent. Reverse blackish green. Colonies<br />

on OA reaching 24–28 mm diam, olive (3F3), granular in concentric<br />

circles, consisting of two kinds of conidiophores (low <strong>and</strong> high), flat,<br />

with flat margin. Reverse black. Colonies on MEA reaching 23–28<br />

mm diam, olive (3E5), deeply furrowed, velvety (sporulating all over)<br />

with undulate, white margin. Reverse brownish green. Colonies on<br />

MEA + 5 % NaCl reaching 28–43 mm diam, olive (3E6), granular<br />

due to profuse sporulation, slightly furrowed with flat, olive-grey<br />

(3F2) margin. Reverse dark green.<br />

Maximum tolerated salt concentration: Only one of three strains<br />

tested (<strong>CBS</strong> 452.71) developed colonies at 17 % NaCl after 14 d,<br />

the other two strains grew until 10 % NaCl.<br />

Cardinal temperatures: For one of three strains (<strong>CBS</strong> 452.71) the<br />

minimum temperature of growth was 4 °C (6 mm diam), for the<br />

other two 10 °C (8–9 mm diam); optimum 25 °C (23–28 mm diam),<br />

maximum 30 °C (only strain <strong>CBS</strong> 452.71 grew 5 mm diam), no<br />

growth at 37 °C.<br />

Specimen examined: Slovenia, from hypersaline water of Sečovlje salterns, coll.<br />

<strong>and</strong> isol. L. Butinar, Dec. 1999, <strong>CBS</strong> H-19732, holotype, culture ex-type EXF-449<br />

= <strong>CBS</strong> 119414.<br />

Habitats <strong>and</strong> distribution: Osmotic environments worldwide.<br />

Differential parameters: Oblong conidia, relatively low degree of<br />

halotolerance.<br />

Strains examined: <strong>CBS</strong> 452.71, EXF-397, EXF-449 (= <strong>CBS</strong> 119414;<br />

ex-type strain).<br />

169


Zalar et al.<br />

Fig. 6. <strong>Cladosporium</strong> dominicanum. Macro- <strong>and</strong> micromorphological characters. A–D. Colony surface grown on PDA (A), OA (B), MEA (C) <strong>and</strong> MEA plus 5 % NaCl (D) of strains<br />

incubated for 14 d at 25 ºC in darkness. E–F. Habit of conidiophores. G. Conidiophore. H–I. Secondary ramoconidia <strong>and</strong> conidia. E–I. All from 7-d-old SNA slide cultures. A, D,<br />

F–H, from EXF-2519; B, C, E from EXF-727; I, EXF-732 (ex-type strain). Scale bars A–D = 10 mm, E = 100 µm, F = 30 µm, G–I = 10 µm.<br />

170


<strong>Cladosporium</strong> sphaerospermum species complex<br />

Fig. 7. <strong>Cladosporium</strong> fusiforme. Macro- <strong>and</strong> micromorphological characters. A–D. Colony surface grown on PDA (A), OA (B), MEA (C) <strong>and</strong> MEA plus 5 % NaCl (D) of strains<br />

incubated for 14 d at 25 ºC in darkness. E–G. Habit of conidiophores. H–I. Ramoconidia <strong>and</strong> conidia. E–I. All from 7-d-old SNA slide cultures. A–H, from EXF-449 (ex-type strain);<br />

I, from <strong>CBS</strong> 452.71. Scale bars A–D = 10 mm, E = 100 µm, F–G = 30 µm, H–I = 10 µm.<br />

www.studiesinmycology.org<br />

171


Zalar et al.<br />

Fig. 8. <strong>Cladosporium</strong> halotolerans. Macro- <strong>and</strong> micromorphological characters. A–D. Colony surface grown on PDA (A), OA (B), MEA (C) <strong>and</strong> MEA plus 5 % NaCl (D) of strains<br />

incubated for 14 d at 25 ºC in darkness. E–H. Habit of conidiophores. I. Conidiophore. J. Succession of secondary ramoconidia. K. Conidia. E–K. All from 7-d-old SNA slide<br />

cultures. A–B, from EXF-572 (ex-type strain); C–D, from EXF-977; E, G, from EXF-972; F, from EXF-564; H, I, K, from EXF-1072; J, from dH 12862. Scale bars A–D = 10 mm,<br />

E = 100 µm, F–G = 50 µm, H = 30 µm, I–K = 10 µm.<br />

<strong>Cladosporium</strong> halotolerans Zalar, de Hoog & Gunde-Cimerman<br />

sp. nov. MycoBank MB492439. Fig. 8.<br />

Etymology: Refers to its halotolerant habit.<br />

Conidiophora erecta, lateralia vel terminalia ex hyphis rectis oriunda; stipes<br />

longitudine variabili, (5–)10–50(–300) × (2–)2.5–3(–5.5) μm, pallide olivaceobrunneus,<br />

levis vel leniter verruculosus, tenuitunicatus, 0–3-septatus, interdum<br />

pluriseptatus, simplex, denticulatus. Conidiorum catenae undique divergentes,<br />

terminales ad 9 conidia continentes. Cellulae conidiogenae indistinctae. Conidia<br />

verrucosa, brunnea vel fusca, unicellularia, plerumque subglobosa vel globosa,<br />

raro breviter ovoidea, utrinque angustata, (2–)3–4(–6) × (2–)2.5–3(–5) μm, long.<br />

: lat. 1.2–1.5; ramoconidia secundaria cylindrica vel quasi globosa, 0(–1)-septata,<br />

(5–)7–12(–37.5) × (2–)2.5–3(–6.5) μm , ad 4 cicatrices terminales ferentia; cicatrices<br />

inspissatae, conspicuae, protuberantes, 0.7–1.0(–1.5) μm diam. Hyphae vagina<br />

polysaccharidica carentes.<br />

172


<strong>Cladosporium</strong> sphaerospermum species complex<br />

Mycelium partly submerged, partly superficial; hyphae without<br />

extracellular polysaccharide-like material. Conidiophores erect,<br />

arising laterally <strong>and</strong> terminally from straight hyphae, stipes of<br />

variable length, (5–)10–50(–300) × (2–)2.5–3(–5.5) μm, pale<br />

olivaceous-brown, smooth to minutely verruculose, thin-walled, 0–<br />

3-septate, unbranched, with pronounced denticles. Conidial chains<br />

branching in all directions, terminal chains with up to 9 conidia.<br />

Conidiogenous cells undifferentiated. Ramoconidia rarely formed.<br />

Conidia verrucose, brown to dark brown, non-septate, usually<br />

subglobose to globose, less often short-ovoid, narrower at both<br />

ends, length : width ratio = 1.2–1.5; (2–)3–4(–6) × (2–)2.5–3(–5)<br />

μm [av. (± SD) 3.5 (± 0.7) × 2.7 (± 0.5)]; secondary ramoconidia<br />

cylindrical to almost spherical, 0–1-septate, (5–)7–12(–37.5) ×<br />

(2–)2.5–3(–6.5) μm [av. (± SD) 10.3 (± 4.8) × 2.9 (± 0.6)], with up<br />

to 4 distal scars. Conidiogenous scars thickened <strong>and</strong> conspicuous,<br />

protuberant, 0.7–1.0(–1.5) μm diam.<br />

Cultural characteristics: Colonies on PDA reaching 27–43 mm diam,<br />

olive (2F5), slightly furrowed, often covered with grey secondary<br />

mycelium, except at the marginal area where only sporulating<br />

structures can be observed. Margin white <strong>and</strong> regular, with<br />

submerged hyphae. Reverse pale green to black. Colonies on OA<br />

reaching 29–40 mm diam, olive (2F6), flat, uniform, granular due<br />

to profuse sporulation <strong>and</strong> fasciculate bundles of conidiophores,<br />

without sterile mycelium. Reverse dark green to black. Colonies on<br />

MEA reaching 18–44 mm diam, highly variable in colour, but mainly<br />

olive (2E5), <strong>and</strong> from flat with regular margin to deeply furrowed<br />

with undulate margin. Colony centre wrinkled with crater-shaped<br />

appearance. Reverse pale to dark green. Colonies on MEA + 5 %<br />

NaCl reaching 24–48 mm diam, olive (3E8), furrowed, velvety, with<br />

more pale, undulate margins. Reverse dark green to black.<br />

Maximum tolerated salt concentration: Only 15 % of tested strains<br />

develop colonies at 20 % NaCl after 7 d, whereas after 14 d all<br />

cultures grow <strong>and</strong> sporulate.<br />

Cardinal temperatures: No growth at 4 °C, optimum 25 °C (18–44<br />

mm diam), maximum 30 °C (6–23 mm diam). No growth at 37 °C.<br />

Specimen examined: Namibia, from hypersaline water of salterns, coll. Nina Gunde-<br />

Cimerman, 1 Sep. 2000, isol. P. Zalar, 1 Oct. 2000, <strong>CBS</strong> H-19734, holotype, culture<br />

ex-type EXF-572 = <strong>CBS</strong> 119416.<br />

Habitats <strong>and</strong> distribution: Hypersaline water in subtropical climates;<br />

indoor environments; Arctic ice; contaminant in lesions of humans<br />

<strong>and</strong> animals; plant phyllosphere; rock.<br />

Literature: Haubold et al. (1998), Meklin et al. (2004).<br />

Differential parameters: Verrucose conidia, short unbranched <strong>and</strong><br />

non-septate conidiophores which arise laterally alongside erect<br />

hyphae.<br />

Strains examined: <strong>CBS</strong> 191.54, <strong>CBS</strong> 573.78, <strong>CBS</strong> 626.82, dH<br />

12862, dH 12991, dH 13911, EXF-228, EXF-380, EXF-565, EXF-<br />

567, EXF-571, EXF-572 (= <strong>CBS</strong> 119416; ex-type strain), EXF-646,<br />

EXF-698, EXF-703, EXF-944, EXF-972, EXF-977, EXF-1072,<br />

EXF-2372.<br />

Notes: <strong>Cladosporium</strong> halotolerans strongly resembles C.<br />

sphaerospermum. Several strains of this species such as dH<br />

12862, dH 12941, <strong>CBS</strong> 191.54 <strong>and</strong> UAMH 7686 have been<br />

isolated sporadically from various indoor habitats in Europe, Brazil<br />

<strong>and</strong> the U.S.A. <strong>and</strong> repeatedly from bathrooms in Slovenia (Table<br />

1). Probably sometimes as uncertain culture contaminations, it<br />

has been isolated from plants (GenBank accession no. L25433),<br />

www.studiesinmycology.org<br />

inner organs of a diseased frog (AY361982) <strong>and</strong> human brain<br />

(Kantarcioglu et al. 2002). <strong>The</strong> presence of C. halotolerans<br />

species in gypsum sediments entrapped in Arctic ice, the fact that<br />

it was repeatedly isolated from hypersaline water <strong>and</strong> possibly its<br />

presence in dolphin skin (see Discussion) suggest that it has a<br />

clear preference for (hyper)osmotic habitats. This is supported by<br />

its ability to grow at 20 % NaCl.<br />

<strong>The</strong> teleomorph of C. halotolerans is predicted to be a<br />

Davidiella species. Strain <strong>CBS</strong> 280.49 was isolated by J.A. von Arx<br />

from teleomorphic material of a fungus labelled as Mycosphaerella<br />

hyperici (Auersw.) Starbäck on Hypericum perforatum in<br />

Switzerl<strong>and</strong>. According to Aptroot (2006) this species may belong<br />

in Davidiella <strong>and</strong> produces a Septoria anamorph. In the original<br />

herbarium specimen, <strong>CBS</strong> H-4867, a Mycosphaerella teleomorph<br />

was present, but no sign of a <strong>Cladosporium</strong> anamorph. We assume<br />

that <strong>CBS</strong> 280.49 was a culture contaminant.<br />

<strong>Cladosporium</strong> langeronii (Fonseca, Leão & Nogueira) Vuill.,<br />

Champ. Paras.: 78. 1931. Fig. 9.<br />

Basionym: Hormodendrum langeronii Fonseca, Leão & Nogueira,<br />

Sci. Med. 5: 563. 1927.<br />

≡ <strong>Cladosporium</strong> langeronii (Fonseca, Leão & Nogueira) Cif., Manuale di<br />

Micologia Medica, ed. 2: 488 (1960), comb. superfl.<br />

Mycelium partly submerged, partly superficial; hyphae sometimes<br />

enveloped in polysaccharide-like material. Conidiophores erect or<br />

ascending, micronematous <strong>and</strong> macronematous, stipes of variable<br />

length, (20–)50–130(–200) × (3–)3.5–4.5(–6.5) μm, dark brown,<br />

rough- <strong>and</strong> thick-walled, regularly septate (cell length 9–22 μm),<br />

arising laterally <strong>and</strong> terminally from submerged or aerial hyphae,<br />

branched. Conidial chains dichotomously branched, up to 6 conidia<br />

in the unbranched parts. Conidiogenous cells undifferentiated,<br />

sometimes seceding <strong>and</strong> forming ramoconidia. Ramoconidia<br />

cylindrical, 0–1 septate, (10–)11–22(–42) × (3–)3.5–4.5(–5) µm,<br />

base broadly truncate, 2–3.5 µm wide, slightly thickened <strong>and</strong><br />

somewhat darkened. Conidia irregularly verruculose to sometimes<br />

loosely verrucose, dark brown, non-septate, usually ovoid, length<br />

: width ratio = 1.3–1.5; conidial size (3–)4–5.5(–8) × (2–)3–4(–5)<br />

μm [av. (± SD) 4.8 (± 1.0) × 3.5 (± 0.6)]; secondary ramoconidia<br />

cylindrical to almost spherical, mostly 0–1(–2)-septate, (5.5–)7.5–<br />

12.5(–35.5) × (2.5–)3–4.5(–5.5) μm [av. (± SD) 10.7 (± 4.7) × 3.6 (±<br />

0.8)], with 2, rarely 3 distal scars. Conidiogenous scars thickened<br />

<strong>and</strong> conspicuous, protuberant, 0.9–1.5(–2.3) μm diam.<br />

Cultural characteristics: Colonies on PDA, OA <strong>and</strong> MEA with<br />

restricted growth, attaining 2.5–4.5, 1.5–7.0 <strong>and</strong> 1.0–5.5 mm<br />

diam, respectively. Colonies flat or heaped (up to 3 mm), dark<br />

green (30F4), with black reverse <strong>and</strong> slightly undulate margin with<br />

immersed mycelium. Sporulating on all media. On MEA + 5 % NaCl<br />

growth is faster, colonies attaining 8.5–12.0 mm diam, sporulating<br />

<strong>and</strong> growing deeply into the agar.<br />

Maximum tolerated salt concentration: All strains develop colonies<br />

at 17 % NaCl after 14 d.<br />

Cardinal temperatures: No growth at 4 °C, optimum / maximum 25<br />

°C (1.0–5.5 mm diam), no growth at 30 °C.<br />

Specimen examined: Brazil, from man ulcero-nodular mycosis of h<strong>and</strong> <strong>and</strong> arm,<br />

1927, coll. <strong>and</strong> isol. da Fonseca, <strong>CBS</strong> H-19737, holotype, culture ex-type <strong>CBS</strong><br />

189.54.<br />

Habitats <strong>and</strong> distribution: Polar ice <strong>and</strong> biomats; conifer wood <strong>and</strong><br />

window frame in Europe; humans; strains originating from nasal<br />

mucus (Buzina et al. 2003) have 100 % sequence homology with<br />

173


Zalar et al.<br />

Fig. 9. <strong>Cladosporium</strong> langeronii. Macro- <strong>and</strong> micromorphological characters. A–D. Colony surface grown on PDA (A), OA (B), MEA (C) <strong>and</strong> MEA plus 5 % NaCl (D) of strains<br />

incubated for 14 d at 25 ºC in darkness. E–F. Habit of conidiophores. G–I. Ramoconidia <strong>and</strong> conidia. E–I. All from 7-d-old SNA slide cultures. A–D, from <strong>CBS</strong> 189.54 (ex-type<br />

strain); E, from <strong>CBS</strong> 109868; F–I, from EXF-999. Scale bars A, C–D = 10 mm, B = 5 mm, E = 100 µm, F = 30 µm, G–I = 10 µm.<br />

the strains studied, as well as with a clone from mycorrhizal roots<br />

(Menkis et al. 2005). <strong>The</strong> species is distributed worldwide, without<br />

any apparent predilection for a particular habitat. <strong>The</strong> strains from<br />

clinical cases probably were culture contaminants.<br />

Literature: da Fonseca et al. (1927a, b).<br />

Differential parameters: Restricted growth; lowest salt halotolerance<br />

taxon of all C. sphaerospermum-like species.<br />

Strains examined: <strong>CBS</strong> 189.54 (ex-type strain), <strong>CBS</strong> 601.84, <strong>CBS</strong><br />

101880, <strong>CBS</strong> 109868, dH 11736, dH 12459 = EXF-999, dH 13833<br />

= EXF-1933.<br />

Notes: De Vries (1952) synonymised the isolate identified as<br />

Hormodendrum langeronii with C. sphaerospermum. Strains of this<br />

species have often been identified as C. cladosporioides (Buzina<br />

et al. 2003, Menkis et al. 2005) although it has slightly longer<br />

conidia.<br />

174


<strong>Cladosporium</strong> sphaerospermum species complex<br />

<strong>Cladosporium</strong> psychrotolerans Zalar, de Hoog & Gunde-<br />

Cimerman, sp. nov. MycoBank MB492428. Fig. 10.<br />

Etymology: Refers to its ability to grow at low temperatures.<br />

Mycelium partim submersum; hyphae vagina polysaccharidica carentes.<br />

Conidiophora erecta vel adscendentia; stipes (10–)50–100(–150) × (3–)3.5–4(–7.5)<br />

μm, olivaceo-brunneus, levis, crassitunicatus, compluries regulariter septatus<br />

(cellulis 10–40 μm longis), identidem dichotome ramosus. Conidiorum catenae<br />

undique divergentes, terminales partes simplices ad 4 conidia continentes. Cellulae<br />

conidiogenae indistinctae. Ramoconidia primaria cylindrica, (18–)19–22(–43) ×<br />

(2.5)3–3.5(–4.5) µm, 0(–1)-septata. Conidia leves vel leniter verruculosa, dilute<br />

brunnea, unicellularia, globosa vel ovoidea, (2.5–)3–4(–4.5) × (2–)2.5–3(–3) μm,<br />

long.: lat. 1.3–1.4; ramoconidia secundaria cylindrica, 0–1(–2)-septata, (5–)8–16(–<br />

36) × (2–)2.5–3(–5) μm, ad 4 cicatrices terminales ferentia; cicatrices inspissatae,<br />

conspicuae, 0.5–2 μm diam.<br />

Mycelium partly superficial partly submerged; hyphae without<br />

extracellular polysaccharide-like material. Conidiophores erect or<br />

ascending, macronematous, stipes (10–)50–100(–150) × (3–)3.5–<br />

4(–7.5) μm, olivaceous-brown, smooth or almost so, thick-walled,<br />

regularly septate (cell length 10–40 μm), arising laterally from<br />

aerial hyphae, repeatedly dichotomously branched. Conidial chains<br />

branching in all directions, up to 4 conidia in the unbranched parts.<br />

Ramoconidia sometimes formed, cylindrical, (18–)19–22(–43) ×<br />

(2.5)3–3.5(–4.5) µm, aseptate, rarely 1-septate, with a broadly<br />

truncate base, up to 2 µm wide, unthickened or slightly thickened,<br />

somewhat darkened-refractive. Conidia smooth to minutely<br />

verruculose, light brown, non-septate, spherical to ovoid, length :<br />

width ratio = 1.3–1.4; conidial size (2.5–)3–4(–4.5) × (2–)2.5–3(–3)<br />

μm [av. (± SD) 3.4 (± 0.5) × 2.5 (± 0.2)]; secondary ramoconidia<br />

cylindrical, 0–1(–2)-septate, (5–)8–16(–36) × (2–)2.5–3(–5) μm<br />

[av. (± SD) 12.7 (± 6.5) × 3.0 (± 0.5)], with up to 4 distal scars.<br />

Conidiogenous scars thickened <strong>and</strong> conspicuous, protuberant,<br />

0.5–2 μm diam.<br />

Cultural characteristics: Colonies on PDA reaching 13–18 mm<br />

diam, velvety, olive (3F4) due to profuse sporulation, flat with<br />

straight margin. Reverse dark green. Colonies on OA reaching 13–<br />

15 mm diam, olive (2F8), of granular appearance due to profuse<br />

sporulation; aerial mycelium sparse. Margin regular. Reverse black.<br />

Colonies on MEA reaching 8–15 mm diam, olive (2F4), velvety,<br />

radially furrowed with undulate white margin. Colonies on MEA<br />

with 5 % NaCl growing faster than on other media, reaching 25–27<br />

mm diam, olive (3E6) <strong>and</strong> granular due to profuse sporulation,<br />

either slightly furrowed or heavily wrinkled with regular or undulate<br />

margin. Reverse dark green.<br />

Maximum tolerated salt concentration: 17 % NaCl after 14 d.<br />

Cardinal temperatures: Minimum at 4 °C (5 mm diam), optimum<br />

<strong>and</strong> maximum at 25 °C (8–15 mm diam).<br />

Specimen examined: Slovenia, from hypersaline water of Sečovlje salterns, coll.<br />

<strong>and</strong> isol. S. Sonjak, May 1999, <strong>CBS</strong> H-19730, holotype, culture ex-type EXF-391<br />

= <strong>CBS</strong> 119412.<br />

Habitats <strong>and</strong> distribution: Hypersaline water in the Mediterranean<br />

basin.<br />

Differential parameters: Growth at 4 °C; maximal NaCl<br />

concentration 17 % NaCl, which differentiates it from other species<br />

with <strong>similar</strong> conidia, like C. sphaerospermum, C. halotolerans <strong>and</strong><br />

C. dominicanum.<br />

Strains examined: EXF-326, EXF-332, EXF-391 (= <strong>CBS</strong> 119412;<br />

ex-type strain), EXF-714.<br />

<strong>Cladosporium</strong> salinae Zalar, de Hoog & Gunde-Cimerman, sp.<br />

nov. MycoBank MB492438. Fig. 11.<br />

Etymology: Refers to salterns (= Latin salinae) as the habitat of<br />

this species.<br />

Mycelium partim submersum; hyphae multa rostra lateralia ferentes, hyphae vagina<br />

polysaccharidica involutae. Conidiophora vix distincta, lateralia vel terminalia ex<br />

hyphis aeriis oriunda; stipes longitudine variabili, (5–)25–50(–60) × (2–)2.5–3(–4)<br />

μm, olivaceo-brunneus, levis vel leniter verruculosus, crassitunicatus, irregulariter<br />

dense septatus (cellulis 6–29 μm longis), simplex, interdum ramosus. Conidiorum<br />

catenae undique divergentes, terminales ad 6 conidia continentes. Cellulae<br />

conidiogenae nonnumquam integratae, in summo sequentiam sympodialem<br />

denticulorum formantes. Conidia levia, interdum leniter verruculosa, dilute brunnea,<br />

unicellularia, plerumque fusiformia, (4.5–)5.5–7.5(–10) × (2–)2.5–3(–3.5) μm,<br />

long. : lat. 1.9–2.4; ramoconidia secundaria cylindrica, 0–1(–2)-septata, (7.5–)9.5–<br />

13.5(–19) × (2.5–)2.5–3.5(–4.5) μm, ad 5 cicatrices terminales ferentia; cicatrices<br />

inspissatae, conspicuae, protuberantes, 0.7–1.8 μm diam.<br />

Mycelium partly superficial partly submerged, with numerous<br />

lateral pegs, consistently enveloped in polysaccharide-like<br />

material. Conidiophores poorly differentiated, micronematous,<br />

stipes (5–)25–50(–60) × (2–)2.5–3(–4) μm, olivaceous-brown,<br />

smooth to often minutely verruculose or irregularly rough-walled,<br />

thick-walled, irregularly densely septate (length of cells 6–29 μm),<br />

arising laterally <strong>and</strong> terminally from aerial hyphae, unbranched,<br />

occasionally branched. Conidial chains branching in all directions,<br />

terminal chains with up to 6 conidia. Conidiogenous cells sometimes<br />

integrated, producing sympodial clusters of pronounced denticles<br />

at their distal ends. Conidia usually smooth, occasionally minutely<br />

verruculose, light brown, aseptate, usually oblong ellipsoidal to<br />

fusiform, length : width ratio = 1.9–2.4; (4.5–)5.5–7.5(–10) × (2–)<br />

2.5–3(–3.5) μm [av. (± SD) 6.7 (± 1.3) × 2.9 (± 0.4)]; secondary<br />

ramoconidia cylindrical, 0–1(–2)-septate, (7.5–)9.5–13.5(–19)<br />

× (2.5–)2.5–3.5(–4.5) μm [av. (± SD) 12.1 (± 3.3) × 3.2 (± 0.6)],<br />

with up to 5 distal scars. Conidiogenous scars thickened <strong>and</strong><br />

conspicuous, protuberant, 0.7–1.8 μm diam.<br />

Cultural characteristics: Colonies on PDA reaching 10–27 mm<br />

diam, granular, olive (2E4) due to profuse sporulation, with white<br />

undulate margin. Aerial mycelium absent. Colonies either heaped<br />

or radially furrowed, in the marginal area growing deeply into the<br />

agar. Reverse dark brown to dark green. Colonies on OA reaching<br />

7–20 mm diam, olive (3E6), of granular appearance due to profuse<br />

sporulation, aerial mycelium present. Margin either undulate or<br />

arachnoid, deeply furrowed. Reverse pale brown to dark green.<br />

Colonies on MEA reaching 8–19 mm diam, velvety, reseda-green<br />

(2E6), heaped. Margin furrowed, growing deeply into the agar.<br />

Colonies on MEA with 5 % NaCl growing much faster than on<br />

other media, reaching 25–38 mm diam, of different colours, mostly<br />

reseda-green (2E6) <strong>and</strong> granulate due to profuse sporulation,<br />

margin olive-yellow (2D6). Reverse yellow to dark green.<br />

Maximum tolerated salt concentration: MEA + 17 % NaCl after 14 d.<br />

Cardinal temperatures: No growth at 4 °C, optimum <strong>and</strong> maximum<br />

temperature at 25 °C (8–19 mm diam), no growth at 30 °C.<br />

Specimen examined: Slovenia, from hypersaline water of Sečovlje salterns, coll.<br />

<strong>and</strong> isol. S. Sonjak, Feb. 1999, <strong>CBS</strong> H-19731, holotype, culture ex-type EXF-335<br />

= <strong>CBS</strong> 119413.<br />

Habitats <strong>and</strong> distribution: Hypersaline water in the Mediterranean<br />

basin.<br />

Differential parameters: Sympodial conidiogenous cells with<br />

pronounced denticles, narrow temperature amplitude.<br />

www.studiesinmycology.org<br />

175


Zalar et al.<br />

Fig. 10. <strong>Cladosporium</strong> psychrotolerans. Macro- <strong>and</strong> micromorphological characters. A–D. Colony surface grown on PDA (A), OA (B), MEA (C) <strong>and</strong> MEA plus 5 % NaCl (D) of<br />

strains incubated for 14 d at 25 ºC in darkness. E–F. Conidiophores. G. Apical part of a conidiophore. H–I. Secondary ramoconidia <strong>and</strong> conidia. E–I. All from 7-d-old SNA slide<br />

cultures. All but C, from EXF-391 (ex-type strain); C, from EXF-714. Scale bars A–D = 10 mm, E = 100 µm, F = 50 µm, G–I = 10 µm.<br />

Strains examined: EXF-322, EXF-335 (= <strong>CBS</strong> 119413; ex-type<br />

strain), EXF-604.<br />

Notes: <strong>Cladosporium</strong> salinae morphologically resembles species<br />

of the <strong>genus</strong> Fusicladium because its conidia are oblong ellipsoidal<br />

to fusiform <strong>and</strong> conidiogenous loci of ramoconidia are placed<br />

closely together. As any other <strong>Cladosporium</strong> species, its conidia<br />

show typical cladosporioid scar structures, however. <strong>Cladosporium</strong><br />

salinae seems to have a separate position within the <strong>genus</strong><br />

<strong>Cladosporium</strong> since it seems to be distantly related to any other<br />

described <strong>Cladosporium</strong> species or currently known species<br />

complex within <strong>Cladosporium</strong>.<br />

176


<strong>Cladosporium</strong> sphaerospermum species complex<br />

Fig. 11. <strong>Cladosporium</strong> salinae. Macro- <strong>and</strong> micromorphological characters. A–D. Colony surface grown on PDA (A), OA (B), MEA (C) <strong>and</strong> MEA plus 5 % NaCl (D) of strains<br />

incubated for 14 d at 25 ºC in darkness. E–F. Habit of conidiophores. G. Conidiophore. H. Detail of apical part of conidiophore. I. Conidia. J. Secondary ramoconidia <strong>and</strong> conidia.<br />

E–J. All from 7-d-old SNA slide cultures. A–D, from EXF-604; E–J, from EXF-335 (ex-type strain). Scale bars A–C = 5 mm, D = 10 mm, E = 100 µm, F = 50 µm, G = 30 µm,<br />

H–J = 10 µm.<br />

<strong>Cladosporium</strong> sphaerospermum Penzig, Michelia 2(8): 473.<br />

1882. Fig. 12.<br />

Mycelium partly submerged, partly superficial; hyphae thick, darkly<br />

pigmented <strong>and</strong> densely septate in submerged mycelium, not<br />

enveloped in polysaccharide-like material. Conidiophores erect or<br />

ascending, micronematous <strong>and</strong> macronematous, stipes of variable<br />

length, (10–)45–130(–300) × (2.5–)3–4(–6) μm, olivaceous-brown,<br />

smooth to minutely verruculose, thick-walled, with relatively<br />

dense septation (cells mostly 4.5–23 long), septa darkened <strong>and</strong><br />

www.studiesinmycology.org<br />

somewhat thickened, arising laterally <strong>and</strong> terminally from immersed<br />

or aerial hyphae, either unbranched or branched. Conidial chains<br />

branching in all directions, up to 6 conidia in the unbranched parts.<br />

Conidiogenous cells not differentiated. Ramoconidia often formed,<br />

cylindrical, (11.5–)20.5–40(–48) × (2.5–)3(–3.5) µm, with up to 5<br />

septa, base broadly truncate, 2 µm wide, slightly thickened <strong>and</strong><br />

somewhat darkened-refractive. Conidia verruculose, brown to<br />

dark brown, non-septate, usually subspherical to spherical, less<br />

often short-ovoid, narrower at both ends, with length : width ratio<br />

= 1.1–1.5; conidial size (2.5–)3–4(–7) × (2–)3–3.5(–4.5) μm [av. (±<br />

177


Zalar et al.<br />

Fig. 12. <strong>Cladosporium</strong> sphaerospermum. Macro- <strong>and</strong> micromorphological characters. A–D. Colony surface grown on PDA (A), OA (B), MEA (C) <strong>and</strong> MEA plus 5 % NaCl (D) of<br />

strains incubated for 14 d at 25 ºC in darkness. E–F. Habit of conidiophores. G–I. Ramoconidia <strong>and</strong> conidia. E–I. All from 7-d-old SNA slide cultures. A, C–D, F–H, from <strong>CBS</strong><br />

193.54 (ex-neotype strain); B, from EXF-738; E, EXF-455; I, EXF-458. Scale bars A–D = 10 mm; E = 100 µm; F = 50 µm; G–I = 10 µm.<br />

SD) 3.8 (± 0.8) × 3.1 (± 0.4)]; secondary ramoconidia cylindrical to<br />

almost spherical, 0–3(–4) septate, (4–)8.5–16(–37.5) × (2–)3–3.5(–<br />

5) μm [av. (± SD) 13.1 (± 6.3) × 3.2 (± 0.5)], with up to 4, rarely up<br />

to 6 distal scars. Conidiogenous scars thickened <strong>and</strong> conspicuous,<br />

protuberant, 0.9–1.1(–1.4) μm diam.<br />

Cultural characteristics: Colonies on PDA reaching 21–44 mm diam,<br />

velvety, olive (2F5) due to profuse sporulation, either with white<br />

<strong>and</strong> regular, or exceptionally undulate margin. Aerial mycelium<br />

sparse. Colonies flat or rarely radially furrowed with elevated<br />

colony centre. Exudates not prominent, some strains release<br />

green soluble pigments into the agar. Reverse blackish blue to pale<br />

green. Growth deep into the agar. Colonies on OA reaching 21–<br />

38 mm diam, olive (2F8), of granular appearance due to profuse<br />

<strong>and</strong> uniform sporulation, almost no aerial mycelium. Margin either<br />

regular or arachnoid, deeply radially furrowed. Reverse black.<br />

Colonies on MEA reaching 15–35 mm diam, velvety, linden-green<br />

(2C5), radially furrowed. Colony centre wrinkled, forming a craterlike<br />

structure; margin furrowed, lighter in colour, consisting of<br />

178


<strong>Cladosporium</strong> sphaerospermum species complex<br />

Fig. 13. <strong>Cladosporium</strong> spinulosum. Macro- <strong>and</strong> micromorphological characters. A–D. Colony surface grown on PDA (A), OA (B), MEA (C) <strong>and</strong> MEA plus 5 % NaCl (D) of strains<br />

incubated for 14 d at 25 ºC in darkness. E. Habit of conidiophores. F–J. Conidiophores. K–L. Conidia (also visible in I–J). E–L. All from 7-d-old SNA slide cultures. A–L, from<br />

EXF-334 (ex-type strain). Scale bars A–D = 10 mm, E = 100 µm, F = 30 µm, G–L = 10 µm.<br />

submerged mycelium. Reverse pale to dark brown. Colonies on<br />

MEA with 5 % NaCl growing faster than on other media, reaching<br />

31–60 mm diam, mainly olive (2D4), either being almost flat or<br />

radially furrowed, with margin of superficial mycelium; sporulation<br />

dense. Reverse ochraceous or dark green.<br />

Maximum tolerated salt concentration: On MEA + 20 % NaCl 89 %<br />

of all strains tested develops colonies after 7 d, 96 % after 14 d.<br />

Cardinal temperatures: No growth at 4 °C, optimum 25 °C (15–35<br />

mm diam), maximum 30 °C (2–15 mm diam). No growth at 37 °C.<br />

Specimen examined: Netherl<strong>and</strong>s, from nail of man, 1949, coll. <strong>and</strong> isol. R.W.<br />

Zappey, <strong>CBS</strong> H-19738, neotype designated here, incorrectly selected by de Vries<br />

(1952) as “lectotype”, culture ex-neotype <strong>CBS</strong> 193.54 = ATCC 11289 = IMI 049637.<br />

Habitats <strong>and</strong> distribution: Hypersaline water in mediterranean<br />

<strong>and</strong> tropics; soil <strong>and</strong> plants in temperate climates; indoor wet<br />

cells; humans. <strong>The</strong> species does not seem to have any particular<br />

preference. Human isolates were probably culture contaminants.<br />

Literature: Penzig (1882), de Vries (1952), Ellis (1971), de Hoog et<br />

al. (2000), Samson et al. (2002).<br />

Diagnostic parameters: Thick-walled, melanised, densely septate<br />

mycelium, almost spherical, verruculose to verrucose terminal<br />

conidia, growth on 20 % NaCl after 7 d.<br />

Strains examined: <strong>CBS</strong> 109.14, <strong>CBS</strong> 122.63, <strong>CBS</strong> 190.54, <strong>CBS</strong><br />

192.54, <strong>CBS</strong> 193.54 (ex-neotype strain), <strong>CBS</strong> 102045, CPC 10944,<br />

EXF-131, EXF-328, EXF-385, EXF-446, EXF-455, EXF-458, EXF-<br />

461, EXF-464, EXF-465, EXF-598, EXF-644, EXF-645, EXF-649,<br />

EXF-715, EXF-738, EXF-739, EXF-781, EXF-962, EXF-965, EXF-<br />

1061, EXF-1726, EXF-1732.<br />

www.studiesinmycology.org<br />

179


Zalar et al.<br />

Fig. 14. <strong>Cladosporium</strong> velox. Macro- <strong>and</strong> micromorphological characters. A–D. Colony surface grown on PDA (A), OA (B), MEA (C) <strong>and</strong> MEA plus 5 % NaCl (D) of strains<br />

incubated for 14 d at 25 ºC in darkness. E–F. Habit of conidiophores. G. Conidiophore. H–I. Secondary ramoconidia <strong>and</strong> conidia. E–I. All from 7-d-old SNA slide cultures. A–D,<br />

G, from <strong>CBS</strong> 119417 (ex-type strain); E–F, H–I, from EXF-466. Scale bars A–D = 10 mm, E = 100 µm, F = 50 µm, G = 30 µm, H–I = 10 µm.<br />

<strong>Cladosporium</strong> spinulosum Zalar, de Hoog & Gunde-Cimerman,<br />

sp. nov. MycoBank MB501099. Fig. 13.<br />

Etymology: Refers to its conspicuously digitate conidia.<br />

Conidiophora erecta vel adscendentia; stipites longitudine variabili, (15–)25–50(–<br />

155) × (2.5–)3–4(–5) μm, olivaceo-brunneus, levis, crassitunicatus, 0–6(–9)-septatus<br />

(cellulis 6–20 μm longis), ex hyphis submersis vel aeriis lateraliter vel terminaliter<br />

oriundus, simplex vel ramosus. Conidiorum catenae undique ramosae, ad 4 conidiis<br />

in partibus linearibus continuis cohaerentibus. Cellulae conidiogenae integratae vel<br />

discretae, acervos distales denticulorum conspicuorum sympodialium proferentes.<br />

Conidia echinulata vel digitata, brunnea vel fusca, continua, vulgo subglobosa vel<br />

globosa, (4.5–)5.5–7(–8) × (3–)4–4.5(–5) μm, long.: lat. = 1.1–1.6, digiti 0.6–1.3<br />

μm longi; ramoconidia secundaria etiam digitata, cylindrica vel subglobosa, 0(–1)-<br />

septata, (6–)6.5–8(–18) × (4–)4.5–5(–5.5) μm, 1–3 cicatrices distales ferentia.<br />

Cicatrices inspissatae, conspicuae, protuberantes, 0.8–1.2 μm diam. Hyphae<br />

nonnumquam polysaccharido circumdatae.<br />

Hyphae sometimes enveloped in polysaccharide-like material.<br />

Conidiophores erect or ascending, stipes of variable length, (15–)<br />

180


<strong>Cladosporium</strong> sphaerospermum species complex<br />

25–50(–155) × (2.5–)3–4(–5) μm, olivaceous-brown, smooth,<br />

sometimes irregularly rough-walled to verrucose near the base,<br />

thick-walled, 0–6(–9)-septate (cells mostly 6–20 μm long), arising<br />

laterally <strong>and</strong> terminally from immersed or aerial hyphae, either<br />

unbranched or branched, somewhat tapering towards the apex.<br />

Conidial chains branching in all directions, up to 4 conidia in the<br />

unbranched parts. Conidiogenous cells sometimes integrated,<br />

producing sympodial clusters of pronounced denticles at their distal<br />

ends. Ramoconidia rarely formed. Conidial wall ornamentation<br />

conspicuously digitate, with up to 1.3 μm long projections having<br />

parallel sides <strong>and</strong> blunt ends. Conidia brown to dark brown,<br />

aseptate, usually subspherical to spherical, length : width ratio =<br />

1.1–1.6; conidial size (4.5–)5.5–7(–8) × (3–)4–4.5(–5) μm [av. (±<br />

SD) 6.2 (± 1.0) × 4.2 (± 0.5)]; secondary ramoconidia ornamented<br />

as conidia, cylindrical to almost spherical, 0(–1)-septate, (6–)6.5–<br />

8(–18) × (4–)4.5–5(–5.5) μm [av. (± SD) 8.6 (± 4.0) × 4.8 (± 0.4)],<br />

with up to 3 distal scars. Conidiogenous scars thickened <strong>and</strong><br />

conspicuous, protuberant, 0.8–1.2 μm diam.<br />

Cultural characteristics: Colonies on PDA reaching 20–30 mm<br />

diam, velvety, dull green (29E4) to dark green (29F6) due to profuse<br />

sporulation, either with white <strong>and</strong> regular, or undulate margin. Aerial<br />

mycelium sparse. Colonies flat or radially furrowed with elevated<br />

colony centre. Growth deep into the agar. Exudates not prominent.<br />

Colonies on OA reaching 20–25 mm diam, dull green (29E4) to dark<br />

green (29F6), sometimes olive (3D4), of granular appearance due<br />

to profuse <strong>and</strong> uniform sporulation; almost without aerial mycelium.<br />

Margin arachnoid. Reverse pale brown to black. Colonies on MEA<br />

reaching 17–28 mm diam, velvety, dull green (29E4) to dark green<br />

(29F6), either flat or radially furrowed. Colony centre wrinkled,<br />

forming a crater-like structure; margin furrowed, paler in colour,<br />

consisting of submerged mycelium only. Reverse pale to dark<br />

green. Colonies on MEA with 5 % NaCl reaching 12–18 mm diam,<br />

of different colours, greenish grey (29D2), greyish green (29D5)<br />

to dark green (29F6); colony appearance variable, mostly either<br />

being almost flat with immersed colony centre or radially furrowed,<br />

with white to dark green margin consisting of superficial mycelium;<br />

sporulation dense. Reverse pale to dark green.<br />

Maximum tolerated salt concentration: On MEA + 17 % NaCl, two<br />

of three strains tested developed colonies after 14 d.<br />

Cardinal temperatures: Growth at 4 °C, optimum <strong>and</strong> maximum at<br />

25 °C (17–28 mm). No growth at 30 °C.<br />

Specimen examined: Slovenia, from hypersaline water of Sečovlje salterns, coll.<br />

<strong>and</strong> isol. S. Sonjak, Feb. 1999, <strong>CBS</strong> H-19796, holotype, culture ex-type EXF-334<br />

= <strong>CBS</strong> 119907.<br />

Habitats <strong>and</strong> distribution: Hypersaline water in temperate climate.<br />

Diagnostic parameters: Conidia <strong>and</strong> ramoconidia with a digitate<br />

ornamentation.<br />

Strains examined: EXF-334 (= <strong>CBS</strong> 119907; ex-type strain), EXF-<br />

382.<br />

Notes: <strong>Cladosporium</strong> spinulosum is a member of the C. herbarum<br />

species complex (Figs 2–4) although its globoid conidia are<br />

reminiscent of C. sphaerospermum. Within <strong>Cladosporium</strong>, the<br />

species is unique in having conspicuously digitate conidia <strong>and</strong><br />

ramoconidia. <strong>The</strong> two strains are differing in the size of conidia.<br />

<strong>The</strong> average size of conidia in EXF-334 is 6.2 (± 0.9) × 4.2 (± 0.5)<br />

μm, <strong>and</strong> in EXF-382 it is 3.9 (± 0.6) × 3.3 (± 0.4) μm.<br />

<strong>Cladosporium</strong> velox Zalar, de Hoog & Gunde-Cimerman, sp. nov.<br />

MycoBank MB492435. Fig. 14.<br />

Etymology: Refers to the quick growth of strains of this species.<br />

Mycelium partim submersum; hyphae vagina polysaccharidica carentes.<br />

Conidiophora erecta, lateralia vel terminalia ex hyphis aeriis oriunda; stipes (10–)<br />

25–150(–250) × (2.5–)3–4(–4.5) μm, olivaceo-brunneus, levis, crassitunicatus, ad<br />

7–septatus (cellulis 10–60 μm longis), identidem dichotome ramosus. Conidiorum<br />

catenae undique divergentes, terminales partes simplices ad 5 conidia continentes.<br />

Cellulae conidiogenae indistinctae. Conidia levia vel leniter verruculosa, dilute<br />

brunnea, unicellularia, ovoidea, (2–)3–4(–5.5) × (1.5–)2–2.5(–3) μm, long. : lat.<br />

1.4–1.7; ramoconidia secundaria cylindrica, 0–1-septata, (3.5–)5.5–19(–42) ×<br />

(2–)2.5–3(–4.5) μm, ad 4(–5) cicatrices terminales ferentia; cicatrices inspissatae,<br />

protuberantes, conspicuae, 0.5–1.5 μm diam.<br />

Mycelium partly superficial partly submerged; hyphae without<br />

extracellular polysaccharide-like material. Conidiophores erect,<br />

stipes (10–)25–150(–250) × (2.5–)3–4(–4.5) μm, slightly attenuated<br />

towards the apex, olivaceous-brown, smooth- <strong>and</strong> thick-walled,<br />

arising terminally <strong>and</strong> laterally from aerial hyphae, dichotomously<br />

branched [up to 5(–7)-septate, cell length 10–60 μm]. Ramoconidia<br />

rarely formed. Conidial chains branching in all directions, terminal<br />

chains with up to 5 conidia. Conidia smooth to very finely<br />

verruculose, pale brown, non-septate, ovoid, length : width ratio<br />

= 1.4–1.7; (2–)3–4(–5.5) × (1.5–)2–2.5(–3) μm [av. (± SD) 3.6 (±<br />

0.6) × 2.3 (± 0.2)]; secondary ramoconidia cylindrical, 0–1-septate,<br />

(3.5–)5.5–19(–42) × (2–)2.5–3(–4.5) μm [av. (± SD) 13.4 (± 10.2)<br />

× 2.8 (± 0.5)], with up to 4(–5) distal scars. Conidiogenous scars<br />

thickened <strong>and</strong> conspicuous, protuberant, 0.5–1.5 μm diam.<br />

Cultural characteristics: Colonies on PDA reaching 35–45 mm<br />

diam, velvety, dark green due to profuse sporulation, on some parts<br />

covered with white sterile mycelium, flat with straight white margin.<br />

Reverse dark green to black. Colonies on OA reaching 30–43 mm<br />

diam, dark green, mycelium submerged, aerial mycelium sparse.<br />

Margin regular. Reverse black. Colonies on MEA reaching 30–42<br />

mm diam, pale green, radially furrowed, with raised, crater-shaped<br />

central part, with white, undulate, submerged margin. Sporulation<br />

poor. Colonies on MEA with 5 % NaCl reaching 35–45 mm diam,<br />

pale green, velvety, flat with regular margin. Reverse pale green.<br />

Sporulation poor.<br />

Maximum tolerated salt concentration: 20 % NaCl after 14 d.<br />

Cardinal temperatures: Minimum at 10 °C (9 mm diam), optimum at<br />

25 °C (30–42 mm diam) <strong>and</strong> maximum at 30 °C (5–18 mm diam).<br />

Specimen examined: India, Charidij, isolated from Bambusa sp., W. Gams, <strong>CBS</strong><br />

H-19735, holotype, culture ex-type <strong>CBS</strong> 119417.<br />

Habitats <strong>and</strong> distribution: Hypersaline water in Slovenia; bamboo,<br />

India.<br />

Strains examined: <strong>CBS</strong> 119417 (ex-type strain), EXF-466, EXF-<br />

471.<br />

ACKNOWLEDGEMENTS<br />

<strong>The</strong> authors are grateful to Walter Gams for his comments on the manuscript<br />

<strong>and</strong> for providing Latin diagnoses, <strong>and</strong> to Konstanze Schubert for contributions to<br />

species descriptions. We thank Kazimir Drašlar <strong>and</strong> Marko Lutar for preparing SEM<br />

illustrations, <strong>and</strong> Kasper Luijsterburg, Špela Štrekelj <strong>and</strong> Barbara Kastelic-Bokal for<br />

their technical assistance. <strong>The</strong> research was partially supported by the European<br />

Union-funded Integrated Infrastructure Initiative grant SYNTHESYS Project NL-<br />

TAF-1070.<br />

www.studiesinmycology.org<br />

181


Zalar et al.<br />

REFERENCES<br />

Aihara M, Tanaka T, Ohta T, Takatori K (2002). Effect of temperature <strong>and</strong> water<br />

activity on the growth of <strong>Cladosporium</strong> sphaerospermum <strong>and</strong> <strong>Cladosporium</strong><br />

cladosporioides. Biocontrol Science 7: 193–196.<br />

Aihara M, Tanaka T, Takatori K (2001). <strong>Cladosporium</strong> as the main fungal contaminant<br />

of locations in dwelling environments. Biocontrol Science 6: 49–52.<br />

Aptroot A (2006). Mycosphaerella <strong>and</strong> its anamorphs: 2. Conspectus of<br />

Mycosphaerella. <strong>CBS</strong> Biodiversity Series 5: 1–231.<br />

Arx JA von (1950). Über die Ascusform von <strong>Cladosporium</strong> herbarum (Pers.) Link.<br />

Sydowia 4: 320–324.<br />

Badillet G, Bièvre C de, Spizajzen S (1982). Isolement de dématiées à partir d’ongles<br />

et de squames. Bulletin de la Société Française de Mycologie 11: 69–72.<br />

Braun U, Crous PW, Dugan F, Groenewald JZ, Hoog GS de (2003). Phylogeny <strong>and</strong><br />

taxonomy of cladosporium-like hyphomycetes, including Davidiella gen. nov.,<br />

the teleomorph of <strong>Cladosporium</strong> s. str. Mycological Progress 2: 3–18.<br />

Butinar L, Sonjak S, Zalar P, Plemenitaš A, Gunde-Cimerman N (2005). Melanized<br />

halophilic fungi are eukaryotic members of microbial communities in hypersaline<br />

waters of solar salterns. Botanica Marina 48: 73–79.<br />

Buzina W, Braun H, Freudenschuss K, Lackner A, Stammberger H (2003). Fungal<br />

biodiversity as found in nasal mucus. Medical Mycology 41: 149–161.<br />

Carbone I, Kohn LM (1999). A method for designing primer sets for speciation<br />

studies in filamentous ascomycetes. Mycologia 91: 553–556.<br />

Clements FE, Shear CL (1931). <strong>The</strong> genera of fungi. New York: H.W. Wilson Co.<br />

Crous PW, Braun U, Groenewald JZ (2007a). Mycosphaerella is polyphyletic.<br />

Studies in Mycology 58: 1-32.<br />

Crous PW, Groenewald JZ, Pongpanich K, Himaman W, Arzanlou M, Wingfield MJ<br />

(2004). Cryptic speciation <strong>and</strong> host specificity among Mycosphaerella spp.<br />

occurring on Australian Acacia species grown as exotics in the tropics. Studies<br />

in Mycology 50: 457–469.<br />

Crous PW, Schubert K, Braun U, Hoog GS de, Hocking AD, Shin H-D, Groenewald<br />

JZ (2007b). Opportunistic, human-pathogenic species in the Herpotrichiellaceae<br />

are phenotypically <strong>similar</strong> to saprobic or phytopathogenic species in the<br />

Venturiaceae. Studies in Mycology 58: 185-217.<br />

Crous PW, Verkley GJM, Groenewald JZ (2006). Eucalyptus microfungi known from<br />

culture. 1. Cladoriella <strong>and</strong> Fulvoflamma genera nova, with notes on some other<br />

poorly known taxa. Studies in Mycology 55: 53–63.<br />

Curtis MD, Gore J, Oliver RP (1994). <strong>The</strong> phylogeny of the tomato leaf mould fungus<br />

<strong>Cladosporium</strong> fulvum syn. Fulvia fulva by analysis of rDNA sequences. Current<br />

Genetics 25: 318–322.<br />

David JC (1997). A contribution to the systematics of <strong>Cladosporium</strong>. Revision of the<br />

fungi previously referred to Heterosporium. Mycological Papers 172: 1–157.<br />

Ellis MB (1971). Dematiaceous Hyphomycetes. Commonwealth Mycological<br />

Institute, Kew.<br />

Fonseca FOOR da, Area Leão AE de, Peñido NJJC (1927a). Mycose de type<br />

ulcéro-nodulaire, semblable à la sporotrichose et produite par Hormodendrum<br />

langeronii. Comptes-rendus des Séances de la Société de Biologie 97: 1772–<br />

1774.<br />

Fonseca FOOR da, Area Leão AE de, Peñido NJJC (1927b). Mycose de type<br />

ulcéro-nodulaire, semlable à la sporotrichose et produite par Hormodendrum<br />

langeronii. Sciencia Medica 5: 563–573.<br />

Gams W, Verkleij GJM, Crous PW (2007). <strong>CBS</strong> Course of Mycology, 5 th ed.<br />

Centraalbureau voor Schimmelcultures, Utrecht, Netherl<strong>and</strong>s.<br />

Gerrits van den Ende AHG, Hoog GS de (1999). Variability <strong>and</strong> molecular diagnostics<br />

of the neurotropic species Cladophialophora bantiana. Studies in Mycology 43:<br />

151–162.<br />

Gunde-Cimerman N, Zalar P, Hoog GS de, Plemenitaš A (2000). Hypersaline<br />

water in salterns – natural ecological niches for halophilic black yeasts. FEMS<br />

Microbiology Ecology 32: 235–240.<br />

Haubold EM, Aronson JF, Cowan DF, McGinnis MR, Cooper CR (1998). Isolation<br />

of fungal rDNA from bottlenose dolphin skin infected with Loboa loboi. Medical<br />

Mycology 36: 263–267.<br />

Herr RA, Tarcha EJ, Taborda PR, Taylor JW, Ajello L, Mendoza L (2001).<br />

Phylogenetic analysis of Lacazia loboi places this previously uncharacterised<br />

pathogen within the dimorphic Onygenales. Journal of Clinical Microbiology<br />

39: 309–314.<br />

Ho MHM, Castañeda RF, Dugan FM, Jong SC (1999). <strong>Cladosporium</strong> <strong>and</strong><br />

Cladophialophora in culture: descriptions <strong>and</strong> an exp<strong>and</strong>ed key. Mycotaxon<br />

72: 115–157.<br />

Hocking AD, Miscamble BF, Pitt J (1994). Water relations of Alternaria alternata,<br />

<strong>Cladosporium</strong> cladosporioides, <strong>Cladosporium</strong> sphaerospermum, Curvularia<br />

lunata <strong>and</strong> Curvularia pallescens. Mycological Research 98: 91–94.<br />

Hoog GS de, Gerrits van den Ende AHG (1998). Molecular diagnostics of clinical<br />

strains of filamentous Basidiomycetes. Mycoses 41: 183–189.<br />

Hoog GS de, Guarro J, Gené J, Figueras MA (2000). Atlas of Clinical Fungi, 2 nd ed.<br />

Centraalbureau voor Schimmelcultures, Utrecht / Universitat Rovira i Virgili,<br />

Reus.<br />

Hoog GS de, Zalar P, Urzì C, Leo F de, Yurlova NA, Sterflinger K (1999). Relationships<br />

of dothideaceous black yeasts <strong>and</strong> meristematic fungi based on 5.8S <strong>and</strong> ITS2<br />

rDNA sequence comparison. Studies in Mycology 43: 31–37.<br />

Jeannmougin F, Thompson JD, Gouy M, Higgins DG, Gibson TJ (1998). Multiple<br />

sequence alignment with ClustalX. Trends in Biochemical Sciences 23: 403–<br />

405.<br />

Kantarcioglu AS, Yucel A, Hoog GS de (2002). Case report. Isolation of <strong>Cladosporium</strong><br />

cladosporioides from cerebrospinal fluid. Mycoses 45: 500–503.<br />

Kirk P, Cannon P, David J, Stalpers J (2001). Ainsworth <strong>and</strong> Bisby’s Dictionary of the<br />

Fungi. 9 th edn. Wallingford, UK. CAB International.<br />

Kornerup A, Wanscher JH (1967). Methuen H<strong>and</strong>book of Colour. 2 nd edn. Methuen<br />

Co., London.<br />

Kumar S, Tamura K, Nei M (2004). MEGA3: integrated software for molecular<br />

evolutionary genetics analysis <strong>and</strong> sequence alignment. Briefings in<br />

Bioinformatics 5: 150–163.<br />

Lacaz CS (1996). Paracoccidioides loboi (Fonseca Filho et Arêa Leão, 1940)<br />

Almeida et Lacaz, 1948–1949. Description of the fungus in Latin. Revista<br />

Instituto Medicina Tropical São Paulo 38: 229–231.<br />

Link HF (1816). Observationes in ordines plantarum naturales III. Magazin der<br />

Gesellschaft Naturforschender Freunde Berlin 7: 37–38.<br />

Masclaux F, Guého E, Hoog GS de, Christen R (1995). Phylogenetic relationships of<br />

human-pathogenic <strong>Cladosporium</strong> (Xylohypha) species inferred from partial LS<br />

rRNA sequences. Journal of Medical <strong>and</strong> Veterinary Mycology 33: 327–338.<br />

McKemy JM & Morgan-Jones G (1991). Studies in the <strong>genus</strong> <strong>Cladosporium</strong> sensu<br />

lato. V. Concerning the type species, <strong>Cladosporium</strong> herbarum. Mycotaxon 42:<br />

307–317.<br />

Meklin T, Haugl<strong>and</strong> RA, Reponen T, Varma M, Lummus Z, Bernstein D, Wymer<br />

LJ, Vesper SJ (2004). Quantitative PCR analysis of house dust can reveal<br />

abnormal mold conditions. Journal of Environmental Monitoring 6: 615–620.<br />

Menkis A, Vasiliauskas R, Taylor AFS, Stenlid J, Finlay R (2005). Fungal<br />

communities in mycorrhizal roots of conifer seedlings in forest nurseries under<br />

different cultivation systems, assessed by morphotyping, direct sequencing<br />

<strong>and</strong> mycelial isolation. Mycorrhiza 16: 33–41.<br />

Moricca S, Ragazzi A, Mitchelson KR (1999). Molecular <strong>and</strong> conventional detection<br />

<strong>and</strong> identification of <strong>Cladosporium</strong> tenuissimum on two-needle pine rust<br />

aeciospores. Canadian Journal of Botany 77: 339–347.<br />

O’Donnell K, Cigelnik E (1997). Two divergent intragenomic rDNA ITS2 types within<br />

a monophyletic lineage of the fungus Fusarium are nonorthologous. Molecular<br />

Phylogenetic Evolution 7: 103–116.<br />

Park HG, Managbanag JR, Stamenova EK, Jong SC (2004). Comparative analysis<br />

of common indoor <strong>Cladosporium</strong> species based on molecular data <strong>and</strong> conidial<br />

characters. Mycotaxon 89: 441–451.<br />

Penzig AJO (1882). Fungi Agromiculi. Michelia 2: 385–503.<br />

Pereira PT, Carvalho MM de, Girio FM, Roseiro JC, Amaral-Collaco MT (2002).<br />

Diversity of microfungi in the phylloplane of plants growing in a Mediterranean<br />

ecosystem. Journal of Basic Microbiology 42: 396–407.<br />

Phillips G, Hudson S, Stewart WK (1992). Microbial growth <strong>and</strong> blockage of subfloor<br />

drains in a renal dialysis centre: a problem highlighted. Journal of Hospital<br />

Infection 21: 193–198.<br />

Pitt JI (1979). <strong>The</strong> Genus Penicillium <strong>and</strong> its teleomorphic states Eupenicillium <strong>and</strong><br />

Talaromyces. Academic Press, London.<br />

Prasil K, Hoog GS de (1988). Variability in <strong>Cladosporium</strong> herbarum. Transactions of<br />

the British Mycological Society 90: 49–54.<br />

Samson RA, Hoekstra ES, Frisvad JC, Filtenborg O (2002). Introduction to Food<strong>and</strong><br />

Airborne Fungi, 6 th ed. Centraalbureau voor Schimmelcultures, Utrecht.<br />

Schoch C, Shoemaker RA, Seifert, KA, Hambleton S, Spatafora JW, Crous PW<br />

(2006). A multigene phylogeny of the Dothideomycetes using four nuclear loci.<br />

Mycologia 98: 1043–1054.<br />

Schubert K, Groenewald JZ, Braun U, Dijksterhuis J, Starink M, Hill CF, Zalar P,<br />

Hoog GS de, Crous PW (2007). Biodiversity in the <strong>Cladosporium</strong> herbarum<br />

(Davidiellaceae, Capnodiales), with st<strong>and</strong>ardisation of methods for<br />

<strong>Cladosporium</strong> taxonomy <strong>and</strong> diagnostics. Studies in Mycology 58: 105–156.<br />

Seifert KA, Nickerson NL, Corlett M, Jackson ED, Louis-Seize G, Davies RJ<br />

(2004). Devriesia, a new hyphomycete <strong>genus</strong> to accommodate heat-resistant,<br />

cladosporium-like fungi. Canadian Journal of Botany 82: 914–926.<br />

Summerbell RC, Cooper E, Bunn U, Jamieson F, Gupta AK (2005). Onychomycosis:<br />

a critical study of techniques <strong>and</strong> criteria confirming the etiologic significance of<br />

nondermatophytes. Medical Mycology 43: 39–59.<br />

Swofford DL (2003). PAUP*. Phylogenetic Analysis Using Parsimony (*<strong>and</strong> Other<br />

Methods). Version 4. Sinauer Associates, Sunderl<strong>and</strong>, Massachusetts.<br />

Trejos A (1954). <strong>Cladosporium</strong> carrionii n. sp. <strong>and</strong> the problem of cladosporia<br />

isolated from chromoblastomycosis. Revista de Biología Tropical 2: 75–112.<br />

Vries GA de (1952). Contribution to the knowledge of the <strong>genus</strong> <strong>Cladosporium</strong> Link<br />

ex Fr. Baarn, 121 pp. Dissertation, University of Utrecht.<br />

Vuillemin P (1931). Les champignons parasites et les mycoses de l’homme.<br />

Encyclopédie Mycologique 2: 1–290.<br />

182


<strong>Cladosporium</strong> sphaerospermum species complex<br />

White TJ, Bruns T, Lee S, Taylor J (1990). Amplification <strong>and</strong> direct sequencing of<br />

fungal ribosomal RNA genes for phylogenetics. In: PCR Protocols: a guide to<br />

methods <strong>and</strong> applications (Innis MA, Gelf<strong>and</strong> DH, Sninsky JJ, White TJ, eds).<br />

Academic Press, San Diego, California: 315–322.<br />

Wirsel SGR, Runge-Froböse C, Ahren DG, Kemen E, Oliver RP, Mendgen,<br />

KW (2002). Four or more species of <strong>Cladosporium</strong> sympatrically colonize<br />

Phragmites australis. Fungal Genetics <strong>and</strong> Biology 25: 99–113.<br />

Yano S, Koyabashi K, Kato K (2002). Intrabronchial lesion due to <strong>Cladosporium</strong><br />

sphaerospermum in a healthy, non-asthmatic woman. Mycoses 46: 348–350.<br />

www.studiesinmycology.org<br />

183


available online at www.studiesinmycology.org<br />

doi:10.3114/sim.2007.58.07<br />

Studies in Mycology 58: 185–217. 2007.<br />

Opportunistic, human-pathogenic species in the Herpotrichiellaceae are<br />

phenotypically <strong>similar</strong> to saprobic or phytopathogenic species in the<br />

Venturiaceae<br />

P.W. Crous 1* , K. Schubert 2 , U. Braun 3 , G.S. de Hoog 1 , A.D. Hocking 4 , H.-D. Shin 5 <strong>and</strong> J.Z. Groenewald 1<br />

1<br />

<strong>CBS</strong> Fungal Biodiversity Centre, P.O. Box 85167, 3508 AD, Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; 2 Botanische Staatssammlung München, Menzinger Strasse 67, D-80638 München,<br />

Germany; 3 Martin-Luther-Universität, Institut für Biologie, Geobotanik und Botanischer Garten, Herbarium, Neuwerk 21, D-06099 Halle, Germany; 4 CSIRO Food Science<br />

Australia, 11 Julius Avenue, North Ryde, NSW 2113, Australia; 5 Division of Environmental Science & Ecological Engineering, Korea University, Seoul 136-701, Korea<br />

*Correspondence: Pedro W. Crous, p.crous@cbs.knaw.nl<br />

Abstract: Although morphologically <strong>similar</strong>, species of Cladophialophora (Herpotrichiellaceae) were shown to be phylogenetically distinct from Pseudocladosporium<br />

(Venturiaceae), which was revealed to be synonymous with the older <strong>genus</strong>, Fusicladium. Other than being associated with human disorders, species of Cladophialophora<br />

were found to also be phytopathogenic, or to occur as saprobes on organic material, or in water, fruit juices, or sports drinks, along with species of Exophiala. Caproventuria<br />

<strong>and</strong> Metacoleroa were confirmed to be synonyms of Venturia, which has Fusicladium (= Pseudocladosporium) anamorphs. Apiosporina, based on A. collinsii, clustered<br />

basal to the Venturia clade, <strong>and</strong> appears to represent a further synonym. Several species with a pseudocladosporium-like morphology in vitro represent a sister clade to the<br />

Venturia clade, <strong>and</strong> are unrelated to Polyscytalum. <strong>The</strong>se taxa are newly described in Fusicladium, which is morphologically close to Anungitea, a heterogeneous <strong>genus</strong> with<br />

unknown phylogenetic affinity. In contrast to the Herpotrichiellaceae, which were shown to produce numerous synanamorphs in culture, species of the Venturiaceae were<br />

morphologically <strong>and</strong> phylogenetically more uniform. Several new species <strong>and</strong> new combinations were introduced in Cladophialophora, Cyphellophora (Herpotrichiellaceae),<br />

Exophiala, Fusicladium, Venturia (Venturiaceae), <strong>and</strong> Cylindrosympodium (incertae sedis).<br />

Taxonomic novelties: Cladophialophora australiensis Crous & A.D. Hocking, sp. nov., Cladophialophora chaetospira (Grove) Crous & Arzanlou, comb. nov., Cladophialophora<br />

hostae Crous, U. Braun & H.D. Shin, sp. nov., Cladophialophora humicola Crous & U. Braun, sp. nov., Cladophialophora potulentorum Crous & A.D. Hocking, sp. nov.,<br />

Cladophialophora scillae (Deighton) Crous, U. Braun & K. Schub., comb. nov., Cladophialophora sylvestris Crous & de Hoog, sp. nov., Cylindrosympodium lauri Crous &<br />

R.F. Castañeda, sp. nov., Cyphellophora hylomeconis Crous, de Hoog & H.D. Shin, sp. nov., Exophiala eucalyptorum Crous, sp. nov., Fusicladium africanum Crous, sp. nov.,<br />

Fusicladium amoenum (R.F. Castañeda & Dugan) Crous, K. Schub. & U. Braun, comb. nov., Fusicladium brevicatenatum (U. Braun & Feiler) Crous, U. Braun & K. Schub.,<br />

comb. nov., Fusicladium fagi Crous & de Hoog, sp. nov., Fusicladium intermedium (Crous & W.B. Kendr.) Crous, comb. nov., Fusicladium matsushimae (U. Braun & C.F. Hill)<br />

Crous, U. Braun & K. Schub., comb. nov., Fusicladium pini Crous & de Hoog, sp. nov., Fusicladium ramoconidii Crous & de Hoog, sp. nov., Fusicladium rhodense Crous & M.J.<br />

Wingf., sp. nov., Venturia hystrioides (Dugan, R.G. Roberts & Hanlin) Crous & U. Braun, comb. nov.<br />

Key words: Anungitea, Anungitopsis, Cladophialophora, Exophiala, Fusicladium, phylogeny, Pseudocladosporium, systematics, Venturia.<br />

Introduction<br />

Species of Cladophialophora Borelli are relatively simple<br />

hyphomycetes with brown hyphae that give rise to branched chains<br />

of pale brown conidia. Phylogenetically they are defined to belong to<br />

the Chaetothyriales (Haase et al. 1999, Untereiner 2000), an order<br />

containing numerous opportunists (de Hoog et al. 2000); teleomorph<br />

relationships are with Capronia Sacc. in the Herpotrichiellaceae.<br />

In several cases cladophialophora-like synanamorphs are found<br />

accompanying black yeasts of the <strong>genus</strong> Exophiala J.W. Carmich.<br />

(de Hoog et al. 1995). Braun & Feiler (1995) placed several saprobic<br />

hyphomycetes in Cladophialophora, <strong>and</strong> described Capronia<br />

hanliniana U. Braun & Feiler as teleomorph of Cladophialophora<br />

brevicatenata U. Braun & Feiler. This work was continued by Dugan<br />

et al. (1995), who described an additional teleomorph, Capronia<br />

hystrioides Dugan, R.G. Roberts & Hanlin for Cladophialophora<br />

hachijoensis (Matsush.) U. Braun & Feiler. Untereiner (1997)<br />

reduced Capronia hystrioides to synonymy with C. hanliniana,<br />

<strong>and</strong> placed them in Venturia Sacc. (Venturiaceae, Pleosporales).<br />

<strong>The</strong> concept of Cladophialophora hachijoensis, which is based<br />

on Phaeoramularia hachijoensis Matsush. (Matsushima 1975) is<br />

confused, however, <strong>and</strong> phylogenetic studies have revealed that<br />

isolates attributed to this name in recent studies, were in fact<br />

representatives of three different species in phylogenetically distinct<br />

genera (Braun et al. 2003). <strong>The</strong> separation of Cladophialophora<br />

with Capronia teleomorphs (Herpotrichiellaceae, Chaetothyriales;<br />

commonly isolated as human pathogens), from predominantly<br />

saprobic or phytopathogenic isolates in the Dothideomycetes<br />

was recognised by Braun (1998). Recently the cactus endophyte<br />

Cladophialophora yegresii de Hoog was reported to be the nearest<br />

neighbour of C. carrionii (Trejos) de Hoog et al., a major agent<br />

of human chromoblastomycosis (de Hoog et al. 2007), so that<br />

the main distinction between the two anamorph genera remains<br />

in their phylogenetic positions. Capronia hanliniana <strong>and</strong> C.<br />

hystrioides were again recognised as distinct species, <strong>and</strong> placed<br />

in a new <strong>genus</strong>, Caproventuria U. Braun (Venturiaceae), while their<br />

anamorphs were accommodated in Pseudocladosporium U. Braun.<br />

Caproventuria was primarily distinguished from Venturia based on<br />

its distinct Pseudocladosporium anamorphs. Recently, Crous et al.<br />

(2007b) introduced a third <strong>genus</strong>, namely Sympoventuria Crous &<br />

Seifert, which produces a sympodiella-like anamorph in culture. To<br />

complicate matters further, Beck et al. (2005) concluded, based on<br />

an ITS DNA phylogeny, that the morphology attributed to the form<br />

genera Spilocaea Fr., Pollaccia E. Bald. & Cif., <strong>and</strong> Fusicladium<br />

Bonord. has evolved several times within Venturia, <strong>and</strong> that a single<br />

anamorph <strong>genus</strong> should be used for Venturia, namely Fusicladium<br />

(see Schubert et al. 2003 for additional generic synonyms).<br />

In their treatment of Venturia anamorphs, Schubert et al.<br />

(2003) excluded Pseudocladosporium, <strong>and</strong> stated that its status<br />

needs to be confirmed along with that of other genera such as<br />

Anungitea B. Sutton, Fusicladium <strong>and</strong> Polyscytalum Riess. In the<br />

study by Beck et al. (2005) an isolate of Caproventuria hystrioides<br />

(Pseudocladosporium sp.) was included to confirm the link to<br />

the Venturiaceae, though this was not well resolved, nor was the<br />

status of the older generic names mentioned above addressed.<br />

185


Crous et al.<br />

<strong>The</strong> aim of the present study, therefore, was to use DNA sequence<br />

comparisons in conjunction with morphology in an attempt to clarify<br />

these generic issues, as well as to determine which morphological<br />

characters could be used to distinguish Pseudocladosporium from<br />

Cladophialophora.<br />

Materials <strong>and</strong> methods<br />

Isolates<br />

Cultures were obtained from the Centraalbureau voor<br />

Schimmelcultures (<strong>CBS</strong>) in Utrecht, the Netherl<strong>and</strong>s, or isolated<br />

from plant material incubated in moist chambers to promote<br />

sporulation. Isolates were cultured on 2 % malt extract plates<br />

(MEA; Gams et al. 2007), by obtaining single conidial colonies as<br />

explained in Crous (2002). Colonies were subcultured onto fresh<br />

MEA, oatmeal agar (OA), potato-dextrose agar (PDA) <strong>and</strong> synthetic<br />

nutrient-poor agar (SNA) (Gams et al. 2007), <strong>and</strong> incubated at 25<br />

°C under continuous near-ultraviolet light to promote sporulation.<br />

DNA extraction, amplification <strong>and</strong> phylogeny<br />

Fungal colonies were established on agar plates, <strong>and</strong> genomic<br />

DNA was isolated following the CTAB-based protocol described<br />

in Gams et al. (2007). <strong>The</strong> primers V9G (de Hoog & Gerrits van<br />

den Ende 1998) <strong>and</strong> LR5 (Vilgalys & Hester 1990) were used to<br />

amplify part (ITS) of the nuclear rDNA operon spanning the 3’ end<br />

of the 18S rRNA gene (SSU), the first internal transcribed spacer<br />

(ITS1), the 5.8S rRNA gene, the second ITS region <strong>and</strong> the 5’<br />

end of the 28S rRNA gene (LSU). Four internal primers, namely<br />

ITS4 (White et al. 1990), LR0R (Rehner & Samuels 1994), LR3R<br />

(www.biology.duke.edu/fungi/mycolab/primers.htm), <strong>and</strong> LR16<br />

(Moncalvo et al. 1993), were used for sequencing to ensure good<br />

quality overlapping sequences were obtained. <strong>The</strong> PCR conditions,<br />

sequence alignment <strong>and</strong> subsequent phylogenetic analysis followed<br />

the methods of Crous et al. (2006a). <strong>The</strong> ITS1, ITS2 <strong>and</strong> 5.8S rRNA<br />

gene were only sequenced for isolates of which these data were<br />

not available. <strong>The</strong> ITS data were not included in the analyses but<br />

deposited in GenBank where applicable. Gaps longer than 10<br />

bases were coded as single events for the phylogenetic analyses;<br />

the remaining gaps were treated as missing data. Sequence data<br />

were deposited in GenBank (Table 1) <strong>and</strong> alignments in TreeBASE<br />

(www.treebase.org).<br />

Taxonomy<br />

Structures were mounted in lactic acid, <strong>and</strong> 30 measurements<br />

(× 1 000 magnification) determined wherever possible, with the<br />

extremes of spore measurements given in parentheses. Colony<br />

colours (surface <strong>and</strong> reverse) were assessed after 2–4 wk on OA<br />

<strong>and</strong> PDA at 25 °C in the dark, using the colour charts of Rayner<br />

(1970). All cultures obtained in this study are maintained in the <strong>CBS</strong><br />

collection (Table 1). Nomenclatural novelties <strong>and</strong> descriptions were<br />

deposited in MycoBank (www.MycoBank.org).<br />

Results<br />

DNA phylogeny<br />

Amplicons of approximately 1 700 bases were obtained for the<br />

isolates listed in Table 1. <strong>The</strong>se sequences were used to obtain<br />

additional sequences from GenBank which were added to the<br />

alignment. <strong>The</strong> manually adjusted LSU alignment contained 116<br />

sequences (including the two outgroup sequences) <strong>and</strong> 1 157<br />

characters including alignment gaps (available in TreeBASE).<br />

Of the 830 characters used in the phylogenetic analysis, 326<br />

were parsimony-informative, 79 were variable <strong>and</strong> parsimonyuninformative,<br />

<strong>and</strong> 425 were constant. Neighbour-joining analyses<br />

using three substitution models on the sequence data yielded trees<br />

with identical topologies to one another. <strong>The</strong> neighbour-joining trees<br />

support the same clades as obtained from the parsimony analysis,<br />

but with a different arrangement at the deep nodes, for example,<br />

the clade containing Protoventuria alpina (Sacc.) M.E. Barr (<strong>CBS</strong><br />

140.83) is placed as sister to the Venturiaceae using parsimony but<br />

basal to the Herpotrichiellaceae using neighbour-joining. Because<br />

of the large number of different strain associations in the Venturia<br />

clade (see the small number of strict consensus branches for this<br />

clade in Fig. 1), only the first 5 000 equally most parsimonious trees<br />

(TL = 1 752 steps; CI = 0.392; RI = 0.849; RC = 0.333) were saved,<br />

one of which is shown in Fig. 1.<br />

Bayesian analysis was conducted on the same aligned LSU<br />

data set using a general time-reversible (GTR) substitution model<br />

with inverse gamma rates <strong>and</strong> dirichlet base frequencies. <strong>The</strong><br />

Markov Chain Monte Carlo (MCMC) analysis of 4 chains started<br />

from a r<strong>and</strong>om tree topology <strong>and</strong> lasted 2 000 000 generations.<br />

Trees were saved each 1 000 generations, resulting in 2 000 saved<br />

trees. Burn-in was set at 500 000 generations after which the<br />

likelihood values were stationary, leaving 1 500 trees from which<br />

the consensus tree (Fig. 2) <strong>and</strong> posterior probabilities (PP’s) were<br />

calculated. <strong>The</strong> average st<strong>and</strong>ard deviation of split frequencies<br />

was 0.06683 at the end of the run. <strong>The</strong> same overall topology as<br />

that observed using parsimony was obtained, with the exception<br />

of the position of Anungitopsis speciosa R.F. Castañeda & W.B.<br />

Kendr., which is placed between the Leotiomycetes <strong>and</strong> the<br />

Sordariomycetes based on the Bayesian analysis. Also, <strong>similar</strong> to<br />

the results obtained using neighbour-joining, the clade containing<br />

Protoventuria alpina (<strong>CBS</strong> 140.83) is placed as sister to the<br />

Herpotrichiellaceae <strong>and</strong> not to the Venturiaceae. <strong>The</strong> phylogenetic<br />

affinity of specific genera or species are discussed below.<br />

Taxonomy<br />

Several collections represented novel members of the<br />

Herpotrichiellaceae <strong>and</strong> Venturiaceae, <strong>and</strong> these are described<br />

below. Taxa that were cladophialophora- or pseudocladosporiumlike,<br />

but that clustered elsewhere, are treated under excluded<br />

species.<br />

Members of Chaetothyriales, Herpotrichiellaceae<br />

Cladophialophora australiensis Crous & A.D. Hocking, sp. nov.<br />

MycoBank MB504525. Fig. 3.<br />

Etymology: Named after its country of origin, Australia.<br />

Cladophialophorae carrionii similis, sed conidiis secundis majoribus, (7–)8–12(–15)<br />

× 3–4 µm.<br />

In vitro: Mycelium consisting of branched, septate, smooth, pale<br />

brown, guttulate, 2–3 µm wide hyphae; hyphal coils not seen.<br />

Conidiophores dimorphic; macroconidiophores mononematous,<br />

subcylindrical, multi-septate, straight to curved, up to 150 µm long<br />

(including conidiogenous cells), <strong>and</strong> 4 µm wide, pale to medium<br />

brown, smooth, guttulate; microconidiophores integrated with<br />

hyphae, which terminate in subcylindrical conidiogenous cells that<br />

give rise to branched chains of conidia; conidiophores (including<br />

186


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

conidiogenous cells) up to 5-septate, 50 µm long, with terminal <strong>and</strong><br />

lateral conidiogenous cells. Conidiogenous cells pale to medium<br />

brown, smooth, guttulate, terminal <strong>and</strong> lateral, subcylindrical,<br />

20–35 × 2–3.5 µm, or reduced to indistinct subtruncate to truncate<br />

loci, scars up to 2 µm wide, mono- to polyblastic, proliferating<br />

sympodially, scars neither darkened, thickened, nor refractive.<br />

Conidia pale to medium brown, guttulate, smooth; ramoconidia<br />

subcylindrical, 0–1-septate, 20–35 × 2–3 µm, hila subtruncate,<br />

inconspicuous, up to 2 µm wide, giving rise to branched chains<br />

of conidia; conidia ellipsoid, pale brown, but becoming dark brown<br />

<strong>and</strong> thick-walled in older cultures, guttulate, tapering towards<br />

subtruncate terminal loci, 0–1-septate, occurring in chains of up to<br />

20 conidia, (7–)8–12(–15) × 3–4 µm (older, dark brown conidia are<br />

ellipsoid, up to 5 µm wide).<br />

Cultural characteristics: Colonies erumpent, somewhat spreading,<br />

margins crenate, feathery, aerial mycelium sparse; colonies on<br />

PDA olivaceous-grey to iron-grey (surface); reverse iron-grey; on<br />

OA <strong>and</strong> SNA olivaceous-grey. Colonies reaching 5 mm diam after 2<br />

wk at 25 °C in the dark; colonies fertile. Not able to grow at 37 °C.<br />

Specimen examined: Australia, isolated from apple juice, Dec. 1986, A.D. Hocking,<br />

holotype <strong>CBS</strong> H-19899, culture ex-type <strong>CBS</strong> 112793 = CPC 1377.<br />

Notes: Cladophialophora australiensis is one of two novel species<br />

of Cladophialophora originally isolated from sports drinks in<br />

Australia. Cladophialophora spp. are commonly associated with<br />

human disorders (Honbo et al. 1984, de Hoog et al. 2000, Levin<br />

et al. 2004), <strong>and</strong> thus their occurrence in sports drinks is cause for<br />

concern. However, none of the new species described here had<br />

the ability to grow at 37 °C, <strong>and</strong> therefore it is not expected that<br />

they could pose a danger to humans. Comparing ITS diversity, the<br />

species shows more than 12 % difference to established pathogens<br />

such as C. carrionii <strong>and</strong> C. bantiana (Sacc.) de Hoog et al.<br />

Cladophialophora chaetospira (Grove) Crous & Arzanlou, comb.<br />

nov. MycoBank MB504526. Fig. 4.<br />

Basionym: Septocylindrium chaetospira Grove, J. Bot. Lond. 24:<br />

199. 1886.<br />

≡ Septonema chaetospira (Grove) S. Hughes, Naturalist, London 840: 9.<br />

1952.<br />

≡ Heteroconium chaetospira (Grove) M.B. Ellis, in Ellis, More Dematiaceous<br />

Hyphomycetes: 64. 1976.<br />

In vitro: Mycelium consisting of branched, septate, smooth,<br />

medium brown hyphae, 2–3.5 µm wide. Conidiophores reduced<br />

to conidiogenous cells, or a single supporting cell, 20–40 × 3–4<br />

µm. Conidiogenous cells subcylindrical, erect, straight to irregularly<br />

curved, medium brown, smooth, 15–30 × 3–4 µm. Conidia in<br />

branched, acropetal chains with up to 30 conidia; subcylindrical to<br />

fusiform, medium brown, smooth, tapering slightly at subtruncate<br />

ends, 1(–3)-septate, thin-walled, becoming slightly constricted at<br />

septa of older conidia, (20–)25–30(–45) × 3–4(–5) µm; conidia<br />

remaining attached in long chains; hila neither thickened, nor<br />

darkened-refractive.<br />

Cultural characteristics: Colonies erumpent, convex, spreading,<br />

with sparse to dense aerial mycelium; margins smooth, undulate;<br />

on PDA iron-grey (surface), margins olivaceous-black; reverse<br />

olivaceous-black; on OA olivaceous-grey in the middle due to fluffy<br />

aerial mycelium, iron-grey in wide outer margin; on SNA olivaceousgrey.<br />

Colonies reaching 12 mm diam after 2 wk on PDA at 25 °C in<br />

the dark. Not able to grow at 37 °C.<br />

Specimens examined: China, Yunnan, Yiliang, isolated from Phyllostachys<br />

bambusoides (Gramineae), decaying bamboo, freshwater, 6 Jul. 2003, L. Cai, <strong>CBS</strong><br />

114747; China, Yunnan, stream in Kunming, isolated from bamboo wood, 15 Jun.<br />

2003, C. Lei, <strong>CBS</strong> 115468. Denmark, isolated from roots of Picea abies (Pinaceae),<br />

isol. by D.S. Malla, <strong>CBS</strong> 491.70. Germany, Schleswig-Holstein, Kiel-Kitzeberg,<br />

isolated from wheat field soil, isol. by W. Gams, <strong>CBS</strong> 514.63 = ATCC 16274 = MUCL<br />

8310.<br />

Notes: Two cultures of Heteroconium chaetospira were originally<br />

deposited as Spadicoides minuta L. Cai, McKenzie & K.D. Hyde (Cai<br />

et al. 2004), but later found to represent Heteroconium chaetospira,<br />

a species commonly found on rotting wood in Europe (Ellis 1976).<br />

<strong>The</strong> <strong>genus</strong> Heteroconium Petr. has in recent years been used<br />

to name leaf spotting fungi with chains of brown, disarticulating<br />

conidia (Crous et al. 2006b), which have phylogenetic affinities<br />

to several orders, obviously being polyphyletic. <strong>The</strong> type species<br />

of Heteroconium, H. citharexyli Petr., is a plant pathogen on<br />

Cytharexylum (Petrak 1949) with hitherto unknown phylogenetic<br />

position. <strong>The</strong> fact that H. chaetospira is linked to the Chaetothyriales,<br />

was rather unexpected. <strong>The</strong> species appears to be <strong>similar</strong> to others<br />

placed in Cladophialophora by having short, lateral conidiogenous<br />

cells, <strong>and</strong> long chains of branched subcylindrical conidia that largely<br />

remain attached. It is, however, quite distinct from other members of<br />

Cladophialophora in having medium brown conidia, <strong>and</strong> in lacking<br />

the ellipsoid conidia observed in several species.<br />

Cladophialophora hostae Crous, U. Braun & H.D. Shin, sp. nov.<br />

MycoBank MB504527. Figs 5–6.<br />

Etymology: Epithet derived from the host <strong>genus</strong>, Hosta.<br />

Cladophialophorae scillae similis, sed conidiophoris in vitro brevioribus et leniter<br />

angustioribus, 10–15 × 1.5–2 µm, conidiis brevioribus, (7–)10–15(–20) µm.<br />

In vivo: Leaf spots amphigenous, subcircular to somewhat angularirregular,<br />

1–5 mm wide, scattered to aggregated, sometimes<br />

confluent, pale to medium brown or with a reddish brown tinge,<br />

later greyish brown, margin indefinite or on the upper leaf surface<br />

with a narrow slightly raised marginal line or very narrow lighter<br />

halo, yellowish, ochraceous to brownish. Caespituli epiphyllous,<br />

punctiform to confluent, dingy greyish brown. Mycelium immersed,<br />

forming fusicladium-like hyphal str<strong>and</strong>s or plates; hyphae septate,<br />

sometimes with constrictions at the septa, thin-walled, pale<br />

olivaceous, 1.5–7 µm wide. Stromata immersed, small, 10–40 µm<br />

diam, composed of swollen hyphal cells, subcircular to somewhat<br />

angular-irregular in outline, 2–8 µm diam, wall somewhat<br />

thickened, brown. Conidiophores in small to moderately large<br />

fascicles, loose, divergent to moderately dense, rarely solitary,<br />

arising from stromatic hyphal aggregations, erumpent, erect,<br />

usually unbranched, rarely branched, straight, subcylindrical to<br />

distinctly geniculate-sinuous, 5–40 × 2–5 µm, 0–6-septate, pale<br />

to medium olivaceous to olivaceous-brown, thin-walled, up to 0.5<br />

µm, smooth. Conidiogenous cells integrated, terminal, 5–15(–20)<br />

µm long, sympodial, conidiogenous loci rather inconspicuous<br />

to subdenticulate, flat-tipped, 1–1.5 µm diam, unthickened or<br />

almost so, not to slightly darkened-refractive. Conidia in simple or<br />

branched chains, narrowly ellipsoid-subcylindrical, 10–15 × 1.5–<br />

3.5 µm, 0–1-septate, subhyaline to pale olivaceous, thin-walled,<br />

smooth, ends truncate or with two denticle-like hila in ramoconidia,<br />

(0.75–)1–1.5(–2) µm diam, unthickened or almost so, at most<br />

slightly darkened-refractive.<br />

In vitro: Mycelium composed of branched, smooth, pale olivaceous<br />

to medium brown hyphae, frequently forming hyphal coils, guttulate,<br />

septa inconspicuous, not constricted, hyphae somewhat irregular<br />

in width, 1–2 µm wide. Conidiophores reduced to conidiogenous<br />

cells, integrated in hyphae, terminal, subcylindrical, pale olivaceous<br />

to pale brown, smooth, 0–1-septate, proliferating sympodially at<br />

www.studiesinmycology.org<br />

187


Crous et al.<br />

10 changes<br />

Athelia epiphylla AY586633<br />

Paullicorticium ansatum AY586693<br />

100<br />

Incertae sedis<br />

Fasciatispora petrakii AY083828<br />

Pidoplitchkoviella terricola AF096197<br />

Anungitopsis speciosa <strong>CBS</strong> 181.95<br />

100<br />

94 100<br />

98<br />

81 97<br />

90<br />

90<br />

87<br />

53<br />

100<br />

94<br />

100<br />

Polyscytalum fecundissimum <strong>CBS</strong> 100506<br />

Phlogicylindrium eucalypti DQ923534<br />

Cyphellophora laciniata <strong>CBS</strong> 190.61<br />

Cladophialophora proteae <strong>CBS</strong> 111667<br />

100<br />

84<br />

82<br />

67<br />

100<br />

Phaeococcomyces catenatus AF050277<br />

Exophiala sp. 3 CPC 12171<br />

Exophiala sp. 3 CPC 12173<br />

Exophiala sp. 3 CPC 12172<br />

Glyphium elatum AF346420<br />

67 100<br />

“<strong>Cladosporium</strong>” adianticola DQ008144<br />

“<strong>Cladosporium</strong>” adianticola DQ008143<br />

Metulocladosporiella musae DQ008162<br />

Metulocladosporiella musicola DQ008159<br />

Thysanorea papuana EU041871<br />

Veronaea botryosa EU041874<br />

Ceramothyrium carniolicum AY004339<br />

Cyphellophora hylomeconis <strong>CBS</strong> 113311<br />

Exophiala eucalyptorum CPC 11261<br />

98<br />

100<br />

100<br />

100<br />

Cladophialophora humicola AF050263<br />

Cladophialophora sylvestris <strong>CBS</strong> 350.83<br />

Cladophialophora hostae CPC 10737<br />

Cladophialophora scillae <strong>CBS</strong> 116461<br />

Exophiala jeanselmei AF050271<br />

Exophiala oligosperma AF050289<br />

Exophiala dermatitidis AF050270<br />

Capronia mansonii AY004338<br />

Capronia munkii AF050250<br />

Ramichloridium mackenziei AF050288<br />

Capronia pilosella AF050254<br />

Exophiala sp. 1 <strong>CBS</strong> 115142<br />

Exophiala sp. 2 <strong>CBS</strong> 115143<br />

53<br />

68<br />

90<br />

Exophiala pisciphila DQ823101<br />

Fonsecaea pedrosoi AF050276<br />

Ramichloridium anceps AF050285<br />

Capronia acutiseta AF050241<br />

Fonsecaea pedrosoi AF050275<br />

Cladophialophora australiensis <strong>CBS</strong> 112793<br />

Cladophialophora potulentorum <strong>CBS</strong> 115144<br />

Cladophialophora potulentorum <strong>CBS</strong> 114772<br />

85 89<br />

98<br />

100<br />

Cladophialophora potulentorum <strong>CBS</strong> 112222<br />

Cladophialophora carrionii AF050262<br />

Phialophora americana AF050283<br />

Phialophora americana AF050280<br />

Cladophialophora chaetospira <strong>CBS</strong> 114747<br />

Cladophialophora chaetospira <strong>CBS</strong> 514.63<br />

Cladophialophora chaetospira <strong>CBS</strong> 491.70<br />

Cladophialophora chaetospira <strong>CBS</strong> 115468<br />

Sordariomycetes,<br />

Xylariomycetidae,<br />

Xylariales<br />

Chaetothyriomycetes, Chaetothyriales, Herpotrichiellaceae<br />

Fig. 1. (Page 188–189). One of 5 000 equally most parsimonious trees obtained from a heuristic search with 100 r<strong>and</strong>om taxon additions of the LSU sequence alignment using<br />

PAUP v. 4.0b10. <strong>The</strong> scale bar shows 10 changes, <strong>and</strong> bootstrap support values from 1 000 replicates are shown at the nodes. Thickened lines indicate the strict consensus<br />

branches <strong>and</strong> ex-type sequences are printed in bold face. <strong>The</strong> tree was rooted to two sequences obtained from GenBank (Athelia epiphylla AY586633 <strong>and</strong> Paullicorticium<br />

ansatum AY586693).<br />

apex via 1–2(–3) flat-tipped, minute, denticle-like loci, 1–1.5 µm<br />

wide, 10–15 × 1.5–2 µm; scars minutely darkened <strong>and</strong> thickened,<br />

but not refractive. Conidia in extremely long chains (–60), simple<br />

or branched, subcylindrical, or narrowly ellipsoid, smooth, pale<br />

olivaceous, 0–1-septate, (7–)10–15(–20) × (1.5–)2(–2.5) µm,<br />

hila truncate, 1–1.5 µm wide, minutely thickened <strong>and</strong> darkenedrefractive.<br />

Cultural characteristics: Colonies on PDA erumpent, spreading,<br />

with smooth, undulate margins <strong>and</strong> dense aerial mycelium; surface<br />

hazel (middle), outer zone isabelline; reverse fuscous-black in<br />

middle, isabelline in outer zone. Colonies reaching 25 mm diam<br />

on SNA, <strong>and</strong> 40 mm diam on PDA after 1 mo at 25 °C in the dark;<br />

colonies fertile.<br />

Specimen examined: Korea, Pyongchang, Hosta plantaginea (Hostaceae), 20 Sep.<br />

2003, H.D. Shin, HAL 2030 F, holotype, culture ex-type SMK 19664, CPC 10737 =<br />

<strong>CBS</strong> 121637, CPC 10738–10739.<br />

Notes: Although this species is morphologically <strong>similar</strong> to<br />

Cladophialophora scillae (Deighton) Crous, U. Braun & K.<br />

Schub. described below in this paper, C. hostae is treated as a<br />

separate taxon due to the differences in the length <strong>and</strong> width of its<br />

conidiophores <strong>and</strong> conidia in vitro, as well as 17 bp differences in<br />

the ITS DNA sequence data <strong>and</strong> a distinct ecology causing leafspots<br />

on a different, unrelated host. Based on disease symptoms<br />

caused on the living host leaves, C. hostae is a very unusual,<br />

unexpected member of the <strong>genus</strong> Cladophialophora. In vivo, the<br />

mycelium forms obvious hyphal str<strong>and</strong>s <strong>and</strong> plates which are<br />

characteristic for Fusicladium species. <strong>The</strong> conidiophores <strong>and</strong><br />

conidia are also fusicladium-like. Nevertheless, this species clusters<br />

within the Herpotrichiellaceae, i.e., it has to be placed in the <strong>genus</strong><br />

Cladophialophora. Biotrophic species like C. hostae <strong>and</strong> C. scillae<br />

without phialidic synanamorphs render the differentiation between<br />

Cladophialophora <strong>and</strong> Fusicladium (incl. Pseudocladosporium)<br />

almost impossible without sequence data. Furthermore, the<br />

188


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

10 changes<br />

Satchmopsis brasiliensis DQ195797<br />

Neofabraea malicorticis AY544662<br />

Protoventuria alpina <strong>CBS</strong> 140.83<br />

Clathrosporium intricatum AY616235<br />

Zeloasperisporium hyphopodioides <strong>CBS</strong> 218.95<br />

96 72<br />

94<br />

99<br />

86<br />

71<br />

65<br />

88<br />

59<br />

71<br />

73<br />

91<br />

66<br />

100<br />

99<br />

81<br />

95<br />

100<br />

100<br />

80<br />

Veronaeopsis simplex EU041877<br />

Fusicladium africanum CPC 12828<br />

Fusicladium africanum CPC 12829<br />

Sympoventuria capensis DQ885906<br />

Sympoventuria capensis DQ885904<br />

Sympoventuria capensis DQ885905<br />

100<br />

Fusicladium amoenum <strong>CBS</strong> 254.95<br />

Fusicladium intermedium <strong>CBS</strong> 110746<br />

Fusicladium rhodense CPC 13156<br />

Fusicladium pini <strong>CBS</strong> 463.82<br />

Fusicladium ramoconidii <strong>CBS</strong> 462.82<br />

90<br />

100<br />

Cylindrosympodium lauri <strong>CBS</strong> 240.95<br />

Venturia fraxini <strong>CBS</strong> 374.55<br />

Venturia macularis <strong>CBS</strong> 477.61<br />

Venturia maculiformis <strong>CBS</strong> 377.53<br />

Metacoleroa dickiei DQ384100<br />

Venturia alpina <strong>CBS</strong> 373.53<br />

Fusicladium fagi <strong>CBS</strong> 621.84<br />

Venturia sp. <strong>CBS</strong> 681.74<br />

Caproventuria hanliniana AF050290<br />

Venturia hystrioides <strong>CBS</strong> 117727<br />

Venturia lonicerae <strong>CBS</strong> 445.54<br />

Venturia chlorospora <strong>CBS</strong> 466.61<br />

Fusicladium catenosporum <strong>CBS</strong> 447.91<br />

Venturia helvetica <strong>CBS</strong> 474.61<br />

Venturia minuta <strong>CBS</strong> 478.61<br />

Venturia polygoni-vivipari <strong>CBS</strong> 114207<br />

Venturia atriseda <strong>CBS</strong> 378.49<br />

Venturia viennotii <strong>CBS</strong> 690.85<br />

Venturia anemones <strong>CBS</strong> 370.55<br />

Venturia aceris <strong>CBS</strong> 372.53<br />

Venturia cephalariae <strong>CBS</strong> 372.55<br />

Venturia tremulae var. tremulae <strong>CBS</strong> 257.38<br />

Venturia ditricha <strong>CBS</strong> 118894<br />

Venturia populina <strong>CBS</strong> 256.38<br />

Venturia atriseda <strong>CBS</strong> 371.55<br />

Venturia inaequalis <strong>CBS</strong> 535.76<br />

Fusicladium pomi <strong>CBS</strong> 309.31<br />

Venturia chlorospora <strong>CBS</strong> 470.61<br />

Fusicladium m<strong>and</strong>shuricum <strong>CBS</strong> 112235<br />

Venturia tremulae var. gr<strong>and</strong>identatae <strong>CBS</strong> 695.85<br />

Venturia tremulae var. populi-albae <strong>CBS</strong> 694.85<br />

Fusicladium radiosum <strong>CBS</strong> 112625<br />

Venturia saliciperda <strong>CBS</strong> 214.27<br />

Venturia saliciperda <strong>CBS</strong> 480.61<br />

Fusicladium convolvularum <strong>CBS</strong> 112706<br />

Fusicladium oleagineum <strong>CBS</strong> 113427<br />

Fusicladium phillyreae <strong>CBS</strong> 113539<br />

Venturia nashicola <strong>CBS</strong> 793.84<br />

Venturia pyrina <strong>CBS</strong> 120825<br />

Venturia pyrina <strong>CBS</strong> 331.65<br />

Venturia aucupariae <strong>CBS</strong> 365.35<br />

Venturia crataegi <strong>CBS</strong> 368.35<br />

Apiosporina collinsii CPC 12229<br />

Fusicladium effusum CPC 4524<br />

Fusicladium effusum CPC 4525<br />

60<br />

91<br />

Venturia carpophila AY849967<br />

Fusicladium carpophilum <strong>CBS</strong> 497.62<br />

Venturia cerasi <strong>CBS</strong> 444.54<br />

Leotiomycetes, Helotiales<br />

Dothideomycetes, Pleosporales, Venturiaceae<br />

Fig 1. (Continued).<br />

morphology of C. hostae in vivo <strong>and</strong> in vitro shows remarkable<br />

differences in conidiophore morphology, i.e., the growth in vivo is<br />

characteristically fusicladium-like (conidiophores macronematous,<br />

long, septate), whereas habit in vitro is rather pseudocladosporiumlike<br />

(conidiophores less developed, usually reduced to conidiogenous<br />

cells, short). However, several Fusicladium species have also been<br />

observed to exibit a Pseudocladosporium growth habit in culture,<br />

suggesting this growth plasticity to be rather common, <strong>and</strong> strongly<br />

influenced by growth conditions.<br />

Cladophialophora humicola Crous & U. Braun, sp. nov.<br />

MycoBank MB504528. Figs 7–8.<br />

www.studiesinmycology.org<br />

Etymology: Named after its ecology, namely occurring in soil.<br />

Cladophialophorae bantianae similis, sed conidiis majoribus, (8–)11–14(–17) ×<br />

(1.5–)2(–2.5) µm, locis conidiogenis et hilis angustioribus, 1–1.5 µm latis.<br />

In vitro: Mycelium composed of branched, smooth, pale<br />

olivaceous to pale brown hyphae, frequently forming hyphal coils,<br />

prominently guttulate, not to slightly constricted at the septa, 1–2<br />

µm wide, cells somewhat uneven in width. Conidiophores solitary,<br />

mostly inconspicuous <strong>and</strong> integrated in hyphae, varying from<br />

inconspicuously truncate lateral loci on hyphal cells, 1–1.5 µm wide,<br />

to occasionally terminal conidiophores, 0–3-septate, subcylindrical,<br />

proliferating sympodially, 10–30 × 1.5–3 µm, pale brown, smooth.<br />

189


Crous et al.<br />

Table 1. Isolates used for sequence analysis.<br />

Anamorph Teleomorph Accession number 1 Host Country Collector GenBank numbers 2<br />

Anungitopsis speciosa <strong>CBS</strong> 181.95*; INIFAT C94/135<br />

(ITS, LSU)<br />

Leaf litter of Buchenavia<br />

capitata Cuba R.F. Castañeda EU035401, EU035401<br />

Cladophialophora australiensis <strong>CBS</strong> 112793*; CPC 1377 Sports drink Australia – EU035402, EU035402<br />

Cladophialophora chaetospira <strong>CBS</strong> 114747 Phyllostachys bambusoides China L. Cai EU035403, EU035403<br />

<strong>CBS</strong> 115468; HKUCC 10147 Bamboo China – EU035404, EU035404<br />

<strong>CBS</strong> 491.70 Roots of Picea abies Denmark – EU035405, EU035405<br />

<strong>CBS</strong> 514.63; ATCC 16274; MUCL 8310 Soil, wheat field Germany – EU035406, EU035406<br />

Cladophialophora hostae CPC 10737* Hosta plantaginea Korea H.D. Shin EU035407, EU035407<br />

Cladophialophora humicola <strong>CBS</strong> 117536*; BBA 65570 Soil, arable Germany Z. Zaspel & H. Nirenberg EU035408, AF050263<br />

Cladophialophora potulentorum <strong>CBS</strong> 112222; CPC 1376; FRR 4946 Sports drink Australia N.J. Charley EU035409, EU035409<br />

<strong>CBS</strong> 114772; CPC 1375; FRR 4947 Sports drink Australia N.J. Charley EU035410, EU035410<br />

<strong>CBS</strong> 115144*; CPC 11048; FRR 3318 Apple juice – – DQ008141, DQ008141<br />

Cladophialophora proteae <strong>CBS</strong> 111667*; CPC 1514 Protea cynaroides South Africa L. Viljoen EU035411, EU035411<br />

Cladophialophora scillae <strong>CBS</strong> 116461* Scilla peruviana New Zeal<strong>and</strong> C.F. Hill EU035412, EU035412<br />

Cladophialophora sylvestris <strong>CBS</strong> 350.83 Pinus sylvestris Netherl<strong>and</strong>s – EU035413, EU035413<br />

“<strong>Cladosporium</strong>” adianticola <strong>CBS</strong> 735.87*; ATCC 200931; INIFAT C87/40 Adiantum sp. Cuba R.F. Castañeda & G. Arnold DQ008125, DQ008144<br />

Cylindrosympodium lauri <strong>CBS</strong> 240.95*; INIFAT C95/3-2 Laurus sp.<br />

Spain, Canary<br />

Isl<strong>and</strong>s R.F. Castañeda EU035414, EU035414<br />

Cyphellophora hylomeconis <strong>CBS</strong> 113311* Helomeco velane Korea H.D. Shin EU035415, EU035415<br />

Cyphellophora laciniata <strong>CBS</strong> 190.61*; ATCC 14166; MUCL 9569 Man, skin Switzerl<strong>and</strong> K.M. Wissel EU035416, EU035416<br />

Exophiala eucalyptorum CPC 11261* Eucalyptus sp. New Zeal<strong>and</strong> J. Stalpers EU035417, EU035417<br />

Exophiala sp. 1 <strong>CBS</strong> 115142; CPC 11044; FRR 5582 Fruit-based drink – – DQ008139, EU035418<br />

Exophiala sp. 2 <strong>CBS</strong> 115143*; CPC 11047; FRR 5599 Bottled spring water – – DQ008140, EU035419<br />

Exophiala sp. 3 CPC 12171 Prunus sp. Canada K.A. Seifert EU035420, EU035420<br />

CPC 12172 Prunus sp. Canada K.A. Seifert EU035421, EU035421<br />

CPC 12173 Prunus sp. Canada K.A. Seifert EU035422, EU035422<br />

Fusicladium africanum CPC 12828* Eucalyptus sp. South Africa P.W. Crous EU035423, EU035423<br />

CPC 12829 Eucalyptus sp. South Africa P.W. Crous EU035424, EU035424<br />

Fusicladium amoenum <strong>CBS</strong> 254.95*; ATCC 200947; CPC 3681; IMI 367525; Leaf litter of Eucalyptus Cuba R.F. Castañeda<br />

INIFAT C94/155; MUCL 39143<br />

gr<strong>and</strong>is<br />

EU035425, EU035425<br />

Fusicladium carpophilum Venturia carpophila <strong>CBS</strong> 497.62; ETH 4568 Prunus sp. Switzerl<strong>and</strong> – EU035426, EU035426<br />

190


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

Fusicladium catenosporum <strong>CBS</strong> 447.91* Salix tri<strong>and</strong>ra Germany H. Butin EU035427, EU035427<br />

Fusicladium convolvularum <strong>CBS</strong> 112706*; CPC 3884; IMI 383037 Convolvulus arvensis New Zeal<strong>and</strong> C.F. Hill AY251082, EU035428<br />

Fusicladium effusum CPC 4524 Carya illinoinensis U.S.A. K. Stevenson AY251084, EU035429<br />

CPC 4525 Carya illinoinensis U.S.A. K. Stevenson AY251085, EU035430<br />

Fusicladium fagi <strong>CBS</strong> 621.84*; ATCC 200937 Fagus sylvatica Netherl<strong>and</strong>s G.S. de Hoog EU035431, EU035431<br />

Fusicladium intermedium <strong>CBS</strong> 110746*; CPC 778; IMI 362702 Eucalyptus sp. Madagascar P.W. Crous EU035432, EU035432<br />

Fusicladium m<strong>and</strong>shuricum Venturia m<strong>and</strong>shurica <strong>CBS</strong> 112235*; CPC 3639 Populus simonii China – EU035433, EU035433<br />

Fusicladium oleagineum <strong>CBS</strong> 113427 Olea europaea New Zeal<strong>and</strong> C.F. Hill EU035434, EU035434<br />

Fusicladium phillyreae <strong>CBS</strong> 113539; UPSC 1329 – Portugal B. d’Oliveira EU035435, EU035435<br />

Fusicladium pini <strong>CBS</strong> 463.82* Pinus sylvestris Netherl<strong>and</strong>s G.S. de Hoog EU035436, EU035436<br />

Fusicladium pomi Venturia inaequalis <strong>CBS</strong> 309.31 – – – EU035437, EU035437<br />

<strong>CBS</strong> 535.76 Sorbus aria Switzerl<strong>and</strong> – EU035460, EU035460<br />

Fusicladium radiosum Venturia tremulae <strong>CBS</strong> 112625; CPC 3638 Populus tremula France – EU035438, EU035438<br />

Fusicladium ramoconidii <strong>CBS</strong> 462.82*; CPC 3679 Pinus sp. Netherl<strong>and</strong>s G.S. de Hoog AY251086, EU035439<br />

Fusicladium rhodense CPC 13156* Ceratonia siliqua Greece P.W. Crous EU035440, EU035440<br />

Polyscytalum fecundissimum <strong>CBS</strong> 100506 Fagus sylvatica Netherl<strong>and</strong>s W. Gams EU035441, EU035441<br />

Zeloasperisporium hyphopodioides <strong>CBS</strong> 218.95*; IMI 367520; INIFAT C94/114; MUCL 39155 Air Cuba R. F. Castañeda EU035442, EU035442<br />

Apiosporina collinsii CPC 12229 Amelanchier alnifolia Canada L.J. Hutchinson EU035443, EU035443<br />

Protoventuria alpina <strong>CBS</strong> 140.83 Arctostaphylos uva-ursi Switzerl<strong>and</strong> – EU035444, EU035444<br />

Sympoventuria capensis <strong>CBS</strong> 120136; CPC 12838 Eucalyptus sp. South Africa P.W. Crous DQ885906, DQ885906<br />

CPC 12839 Eucalyptus sp. South Africa P.W. Crous DQ885905, DQ885905<br />

CPC 12840 Eucalyptus sp. South Africa P.W. Crous DQ885904, DQ885904<br />

Venturia aceris <strong>CBS</strong> 372.53 Acer pseudoplatanus Switzerl<strong>and</strong> – EU035445, EU035445<br />

Venturia alpina <strong>CBS</strong> 373.53 Arctostaphylos alpina Switzerl<strong>and</strong> – EU035446, EU035446<br />

Venturia anemones <strong>CBS</strong> 370.55; IMI 163998 Anemone alpina France – EU035447, EU035447<br />

Venturia atriseda <strong>CBS</strong> 371.55 Gentiana punctata Switzerl<strong>and</strong> – EU035448, EU035448<br />

<strong>CBS</strong> 378.49 Gentiana lutea Switzerl<strong>and</strong> J.A. von Arx EU035449, EU035449<br />

Venturia aucupariae <strong>CBS</strong> 365.35; IMI 163987 Sorbus aucuparia Germany – EU035450, EU035450<br />

Venturia cephalariae <strong>CBS</strong> 372.55 Cephalaria alpina Switzerl<strong>and</strong> – EU035451, EU035451<br />

Venturia cerasi <strong>CBS</strong> 444.54; ATCC 12119; IMI 163988 Prunus cerasus Germany – EU035452, EU035452<br />

Venturia chlorospora <strong>CBS</strong> 466.61; ETH 543 Salix caesia Switzerl<strong>and</strong> J. Nüesch EU035453, EU035453<br />

<strong>CBS</strong> 470.61 Salix daphnoides France J. Nüesch EU035454, EU035454<br />

Venturia crataegi <strong>CBS</strong> 368.35 Crataegus sp. Germany – EU035455, EU035455<br />

www.studiesinmycology.org<br />

191


Crous et al.<br />

Table 1. (Continued).<br />

Anamorph Teleomorph Accession number 1 Host Country Collector GenBank numbers 2<br />

Venturia ditricha <strong>CBS</strong> 118894 Betula pubescens var.<br />

tortuosa<br />

(ITS, LSU)<br />

Finl<strong>and</strong> – EU035456, EU035456<br />

Venturia fraxini <strong>CBS</strong> 374.55 Fraxinus excelsior Switzerl<strong>and</strong> – EU035457, EU035457<br />

Venturia helvetica <strong>CBS</strong> 474.61; ETH 2571; IMI 163990 Salix helvetica Switzerl<strong>and</strong> J. Nüesch EU035458, EU035458<br />

Venturia hystrioides <strong>CBS</strong> 117727*; ATCC 96019; CPC 5391 Prunus avium cv. Bing U.S.A. R.G. Roberts EU035459, EU035459<br />

Venturia lonicerae <strong>CBS</strong> 445.54; IMI 163997 Lonicera coerulea Switzerl<strong>and</strong> – EU035461, EU035461<br />

Venturia macularis <strong>CBS</strong> 477.61; ETH 2831 Populus tremula France – EU035462, EU035462<br />

Venturia maculiformis <strong>CBS</strong> 377.53 Epilobium montanum France – EU035463, EU035463<br />

Venturia minuta <strong>CBS</strong> 478.61; ETH 523; IMI 163991 Salix nigricans Switzerl<strong>and</strong> J. Nüesch EU035464, EU035464<br />

Venturia nashicola <strong>CBS</strong> 793.84 Pyrus serotina Japan – EU035465, EU035465<br />

Venturia polygoni-vivipari <strong>CBS</strong> 114207; UPSC 2754 Polygonum viviparum Norway K. & L. Holm EU035466, EU035466<br />

Venturia populina <strong>CBS</strong> 256.38; IMI 163996 Populus canadensis Italy – EU035467, EU035467<br />

Venturia pyrina <strong>CBS</strong> 120825 Pyrus communis Brazil – EU035468, EU035468<br />

<strong>CBS</strong> 331.65 Pyrus sp. – – EU035469, EU035469<br />

Venturia saliciperda <strong>CBS</strong> 214.27; IMI 163993 – – – EU035470, EU035470<br />

<strong>CBS</strong> 480.61; ETH 2836 Salix cordata Switzerl<strong>and</strong> – EU035471, EU035471<br />

Venturia sp. <strong>CBS</strong> 681.74 Cedrus atlantica France W. Gams EU035472, EU035472<br />

Venturia tremulae var. gr<strong>and</strong>identatae <strong>CBS</strong> 695.85 Populus tremuloides Canada – EU035473, EU035473<br />

Venturia tremulae var. populi-albae <strong>CBS</strong> 694.85 Populus alba France – EU035474, EU035474<br />

Venturia tremulae var. tremulae <strong>CBS</strong> 257.38 Populus tremula Italy – EU035475, EU035475<br />

Venturia viennotii <strong>CBS</strong> 690.85 Populus tremula France – EU035476, EU035476<br />

1 ATCC: American Type Culture Collection, Virginia, U.S.A.; BBA: Biologische Bundesanstalt für L<strong>and</strong>- und Forstwirtschaft, Berlin-Dahlem, Germany; <strong>CBS</strong>: Centraalbureau voor Schimmelcultures, Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; CPC: Culture collection of Pedro<br />

Crous, housed at <strong>CBS</strong>; ETH: Eidgenössische Technische Hochschule, Institute for Special Botany, Zürich, Switzerl<strong>and</strong>; FRR: Division of Food Research, CSIRO, North Ryde, N.S.W., Australia; HKUCC: <strong>The</strong> University of Hong Kong Culture Collection, Dept.<br />

of Ecology <strong>and</strong> Biodiversity, University of Hong Kong, Pokfulam Road, China; IMI: International Mycological Institute, CABI-Bioscience, Egham, Bakeham Lane, U.K.; INIFAT: Alex<strong>and</strong>er Humboldt Institute for Basic Research in Tropical Agriculture, Ciudad<br />

de La Habana, Cuba; MUCL: Mycotheque de l’ Université Catholique de Louvain, Louvain-la-Neuve, Belgium; UPSC: Uppsala University Culture Collection of Fungi, Museum of Evolution, Botany Section, Evolutionary Biology Centre, Uppsala, Sweden.<br />

2 ITS: internal transcribed spacer regions, LSU: partial 28S rDNA sequence.<br />

*Ex-type cultures.<br />

192


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

0.1 expected changes per site<br />

Athelia epiphylla AY586633<br />

Paullicorticium ansatum AY586693<br />

0.92 0.98 Neofabraea malicorticis AY544662<br />

Protoventuria alpina <strong>CBS</strong> 140.83<br />

0.94 Clathrosporium intricatum AY616235<br />

Leotiomycetes, Helotiales<br />

Satchmopsis brasiliensis DQ195797<br />

1.00<br />

Anungitopsis speciosa <strong>CBS</strong> 181.95 Incertae sedis<br />

0.98 1.00 Fasciatispora petrakii AY083828<br />

Sordariomycetes,<br />

Pidoplitchkoviella terricola AF096197<br />

0.98 Polyscytalum fecundissimum <strong>CBS</strong> 100506 Xylariomycetidae,<br />

1.00 Phlogicylindrium eucalypti DQ923534 Xylariales<br />

1.00<br />

0.94<br />

Cladophialophora humicola AF050263<br />

0.84 Cladophialophora sylvestris <strong>CBS</strong> 350.83<br />

1.00 Cladophialophora hostae CPC 10737<br />

Cladophialophora scillae <strong>CBS</strong> 116461<br />

0.50 Cladophialophora proteae <strong>CBS</strong> 111667<br />

Phaeococcomyces catenatus AF050277<br />

1.00 Glyphium elatum AF346420<br />

1.00 Exophiala sp. 3 CPC 12171<br />

1.00 Exophiala sp. 3 CPC 12173<br />

1.00 Exophiala sp. 3 CPC 12172<br />

“<strong>Cladosporium</strong>” adianticola DQ008144<br />

1.00<br />

1.00<br />

1.00<br />

0.85<br />

0.56<br />

1.00<br />

1.00<br />

1.00<br />

“<strong>Cladosporium</strong>” adianticola DQ008143<br />

Metulocladosporiella musae DQ008162<br />

Metulocladosporiella musicola DQ008159<br />

Thysanorea papuana EU041871<br />

Veronaea botryosa EU041874<br />

Exophiala jeanselmei AF050271<br />

Exophiala oligosperma AF050289<br />

Ramichloridium mackenziei AF050288<br />

Ramichloridium anceps AF050285<br />

Capronia pilosella AF050254<br />

Capronia acutiseta AF050241<br />

Exophiala dermatitidis AF050270<br />

Capronia mansonii AY004338<br />

Capronia munkii AF050250<br />

0.71<br />

1.00<br />

0.81 Exophiala pisciphila DQ823101<br />

0.56 Fonsecaea pedrosoi AF050276<br />

0.98 Exophiala sp. 1 <strong>CBS</strong> 115142<br />

0.60 Exophiala sp. 2 <strong>CBS</strong> 115143<br />

0.80 Cyphellophora laciniata <strong>CBS</strong> 190.61<br />

Ceramothyrium carniolicum AY004339<br />

1.00 Cyphellophora hylomeconis <strong>CBS</strong> 113311<br />

1.00 Exophiala eucalyptorum CPC 11261<br />

Fonsecaea pedrosoi AF050275<br />

0.77 0.68<br />

Cladophialophora australiensis <strong>CBS</strong> 112793<br />

Cladophialophora potulentorum <strong>CBS</strong> 115144<br />

1.00 Cladophialophora potulentorum <strong>CBS</strong> 114772<br />

1.00 Cladophialophora potulentorum <strong>CBS</strong> 112222<br />

1.00<br />

0.95<br />

1.00<br />

1.00<br />

1.00<br />

Cladophialophora carrionii AF050262<br />

Phialophora americana AF050283<br />

Phialophora americana AF050280<br />

Cladophialophora chaetospira <strong>CBS</strong> 114747<br />

Cladophialophora chaetospira <strong>CBS</strong> 514.63<br />

Cladophialophora chaetospira <strong>CBS</strong> 491.70<br />

Cladophialophora chaetospira <strong>CBS</strong> 115468<br />

Fig. 2. (Page 193–194). Consensus phylogram (50 % majority rule) of 1 500 trees resulting from a Bayesian analysis of the LSU sequence alignment using MrBayes v. 3.1.2.<br />

Bayesian posterior probabilities are indicated at the nodes. Ex-type sequences are printed in bold face. <strong>The</strong> tree was rooted to two sequences obtained from GenBank (Athelia<br />

epiphylla AY586633 <strong>and</strong> Paullicorticium ansatum AY586693).<br />

Chaetothyriomycetes, Chaetothyriales, Herpotrichiellaceae<br />

Conidiogenous cells integrated, inconspicuous, truncate, lateral<br />

loci 1–1.5 µm wide, or conidiogenous cells subcylindrical with 1–3<br />

sympodial loci (which appear as minute lateral denticles), 7–17 ×<br />

1.5–2 µm; scars inconspicuous, neither darkened, refractive nor<br />

thickened. Conidia in short chains of up to 10, simple or branched,<br />

subcylindrical to narrowly ellipsoid, 0–1-septate, (8–)11–14(–17)<br />

× (1.5–)2(–2.5) µm, pale olivaceous to olivaceous-brown or pale<br />

brown, smooth, hila truncate, 1–1.5 µm wide, unthickened, neither<br />

darkened, nor refractive.<br />

Cultural characteristics: Colonies erumpent, spreading, with<br />

uneven, feathery margins <strong>and</strong> dense aerial mycelium on PDA;<br />

pale olivaceous-grey in the middle, becoming olivaceous-grey<br />

in the outer zone (surface); reverse olivaceous-black, with greyolivaceous<br />

margins. Colonies reaching 7 mm diam after 2 wk at 25<br />

°C in the dark; colonies fertile.<br />

Specimen examined: Germany, Br<strong>and</strong>enburg, Müncheberg, from soil, Zaspel,<br />

Zalf & H. Nirenberg, holotype <strong>CBS</strong> H-19902, culture ex-type BBA 65570 = <strong>CBS</strong><br />

117536.<br />

Notes: Phylogenetically Cladophialophora humicola is closely<br />

related to C. sylvestris Crous & de Hoog (see below). Morphologically<br />

the two species can be distinguished in that C. humicola lacks<br />

ramoconidia, <strong>and</strong> has 1-septate conidia, while those of C. sylvestris<br />

are 0–3-septate.<br />

Cladophialophora potulentorum Crous & A.D. Hocking, sp. nov.<br />

MycoBank MB504529. Figs 9–10.<br />

Etymology: Refers to its presence in fruit juices <strong>and</strong> sports drinks.<br />

Cladophialophorae carrionii similis, sed conidiis secundis majoribus, (6–)8–10(–13)<br />

× 2–3 µm.<br />

In vitro: Mycelium consisting of branched, septate, smooth,<br />

pale brown, guttulate, 1.5–2.5 µm wide hyphae. Conidiophores<br />

solitary, macronematous, well distinguishable under the dissecting<br />

microscope from aerial mycelium, pale to medium brown,<br />

subcylindrical, straight to somewhat curved, erect, with apical<br />

apparatus appearing as a tuft due to extremely long conidial<br />

www.studiesinmycology.org<br />

193


Crous et al.<br />

0.1 expected changes per site<br />

Cylindrosympodium lauri <strong>CBS</strong> 240.95<br />

1.00 Fusicladium amoenum <strong>CBS</strong> 254.95<br />

1.00 Fusicladium intermedium <strong>CBS</strong> 110746<br />

1.00<br />

Fusicladium rhodense CPC 13156<br />

Fusicladium pini <strong>CBS</strong> 463.82<br />

1.00 1.00 1.00 Fusicladium ramoconidii <strong>CBS</strong> 462.82<br />

1.00<br />

Veronaeopsis simplex EU041877<br />

Fusicladium africanum CPC 12828<br />

Fusicladium africanum CPC 12829<br />

0.85<br />

0.95<br />

0.78<br />

0.56<br />

0.71<br />

1.00<br />

0.69<br />

0.79<br />

0.94<br />

0.56<br />

0.97<br />

0.91<br />

1.00<br />

0.97<br />

1.00<br />

1.00<br />

0.68<br />

0.55<br />

0.65<br />

0.61<br />

0.92<br />

0.94<br />

0.72<br />

0.95<br />

Sympoventuria capensis DQ885906<br />

Sympoventuria capensis DQ885904<br />

Sympoventuria capensis DQ885905<br />

Zeloasperisporium hyphopodioides <strong>CBS</strong> 218.95<br />

Metacoleroa dickiei DQ384100<br />

Venturia alpina <strong>CBS</strong> 373.53<br />

Venturia fraxini <strong>CBS</strong> 374.55<br />

Venturia macularis <strong>CBS</strong> 477.61<br />

Venturia maculiformis <strong>CBS</strong> 377.53<br />

Apiosporina collinsii CPC 12229<br />

Fusicladium fagi <strong>CBS</strong> 621.84<br />

Venturia sp. <strong>CBS</strong> 681.74<br />

Caproventuria hanliniana AF050290<br />

Venturia hystrioides <strong>CBS</strong> 117727<br />

Venturia viennotii <strong>CBS</strong> 690.85<br />

Venturia inaequalis <strong>CBS</strong> 535.76<br />

Venturia saliciperda <strong>CBS</strong> 214.27<br />

Fusicladium radiosum <strong>CBS</strong> 112625<br />

Venturia saliciperda <strong>CBS</strong> 480.61<br />

Venturia tremulae var. populi-albae <strong>CBS</strong> 694.85<br />

Venturiatremulae var. gr<strong>and</strong>identatae <strong>CBS</strong> 695.85<br />

Fusicladium m<strong>and</strong>shuricum <strong>CBS</strong> 112235<br />

Fusicladium pomi <strong>CBS</strong> 309.31<br />

Venturia atriseda <strong>CBS</strong> 371.55<br />

Venturia populina <strong>CBS</strong> 256.38<br />

Venturia chlorospora <strong>CBS</strong> 470.61<br />

Venturia ditricha <strong>CBS</strong> 118894<br />

Venturia tremulae var. tremulae <strong>CBS</strong> 257.38<br />

Fusicladium catenosporum <strong>CBS</strong> 447.91<br />

Venturia helvetica <strong>CBS</strong> 474.61<br />

Venturia minuta <strong>CBS</strong> 478.61<br />

Venturia chlorospora <strong>CBS</strong> 466.61<br />

Venturia polygoni-vivipari <strong>CBS</strong> 114207<br />

Venturia lonicerae <strong>CBS</strong> 445.54<br />

Venturia carpophila AY849967<br />

Fusicladium carpophilum <strong>CBS</strong> 497.62<br />

Venturia cerasi <strong>CBS</strong> 444.54<br />

Venturia atriseda <strong>CBS</strong> 378.49<br />

Venturia anemones <strong>CBS</strong> 370.55<br />

Venturia aceris <strong>CBS</strong> 372.53<br />

Venturia cephalariae <strong>CBS</strong> 372.55<br />

Fusicladium convolvularum <strong>CBS</strong> 112706<br />

Venturia aucupariae <strong>CBS</strong> 365.35<br />

Venturia crataegi <strong>CBS</strong> 368.35<br />

Venturia nashicola <strong>CBS</strong> 793.84<br />

Venturia pyrina <strong>CBS</strong> 120825<br />

Venturia pyrina <strong>CBS</strong> 331.65<br />

Fusicladium effusum CPC 4524<br />

Fusicladium effusum CPC 4525<br />

Fusicladium oleagineum <strong>CBS</strong>113427<br />

Fusicladium phillyreae <strong>CBS</strong> 113539<br />

Dothideomycetes, Pleosporales, Venturiaceae<br />

Fig. 2. (Continued).<br />

Fig. 3. Cladophialophora australiensis (<strong>CBS</strong> 112793). A. Conidiophore. B–C. Subcylindrical ramoconidia, <strong>and</strong> ellipsoid conidia. Scale bar = 10 µm.<br />

194


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

Fig. 4. Cladophialophora chaetospira (<strong>CBS</strong> 114747). A–C. Hyphae giving rise to conidiophores with catenulate conidia. D–F. Conidia become up to 3-septate, frequently<br />

remaining attached in chains. Scale bars = 10 µm.<br />

Fig. 5. Cladophialophora hostae (CPC 10737). A–B. Conidiogenous loci (arrows). C. Hyphal coil. D–F. Branched conidial chains. G–H. Conidia. Scale bar = 10 µm.<br />

www.studiesinmycology.org<br />

195


Crous et al.<br />

Fig. 6. Cladophialophora hostae (CPC 10737). Branched conidial chains with<br />

ramoconidia <strong>and</strong> conidia. Scale bar = 10 µm.<br />

Fig. 7. Cladophialophora humicola (<strong>CBS</strong> 117536). Conidiophore with branched<br />

conidial chains. Scale bar = 10 µm.<br />

Fig. 8. Cladophialophora humicola (<strong>CBS</strong> 117536). A. Hyphal coil. B. Conidiophore. C–F. Conidial chains with ramoconidia <strong>and</strong> conidia. Scale bar = 10 µm.<br />

196


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

Fig. 9. Cladophialophora potulentorum (<strong>CBS</strong> 115144). A. Colony on PDA. B. Conidiophore. C–D. Conidial chains. E–F. Ramoconidia <strong>and</strong> conidia. Scale bar = 10 µm.<br />

chains; conidiophores up to 5-septate, <strong>and</strong> 100 µm tall (excluding<br />

conidiogenous cells). Conidiogenous cells pale brown, smooth,<br />

terminal <strong>and</strong> lateral, subcylindrical, tapering towards subtruncate<br />

to truncate loci, 1 µm wide, somewhat darkened, thickened, but not<br />

refractive, loci appearing subdenticulate on lateral conidiogenous<br />

cells, mono- to polyblastic, proliferating sympodially, 10–35 × 1.5–2<br />

µm. Conidia pale brown, smooth, guttulate, occurring in branched<br />

chains of up to 60; hila somewhat darkened <strong>and</strong> thickened, but not<br />

refractive, 0.5 µm wide; ramoconidia subcylindrical, 0–1-septate,<br />

15–17(–20) × 2.5–3 µm; conidia ellipsoid, (6–)8–10(–13) × 2–3<br />

µm.<br />

Cultural characteristics: Colonies erumpent, spreading, with smooth<br />

margins <strong>and</strong> dense aerial mycelium on PDA, olivaceous-grey<br />

(surface), with a thin, olivaceous-black margin; reverse olivaceousblack;<br />

on OA olivaceous-grey (surface) with a wide olivaceousblack<br />

margin. Colonies reaching 25–30 mm diam after 1 mo at 25<br />

°C in the dark; colonies fertile, also sporulating in the agar. Not able<br />

to grow at 37 °C.<br />

Specimens examined: Australia, isolated from apple juice drink, Dec. 1986, A.D.<br />

Hocking, holotype <strong>CBS</strong> H-19901, culture ex-type <strong>CBS</strong> 115144 = CPC 11048;<br />

Australia, isolated from sports drink, Feb. 1996, A.D. Hocking, <strong>CBS</strong> 114772 = CPC<br />

1375 = FRR 4947; Australia, isolated from sports drink, Feb. 1996, A.D. Hocking,<br />

<strong>CBS</strong> 112222 = FRR 4946.<br />

Fig. 10. Cladophialophora potulentorum (<strong>CBS</strong> 115144). Conidiophores with chains<br />

of ramoconidia <strong>and</strong> conidia. Scale bar = 10 µm.<br />

www.studiesinmycology.org<br />

Notes: Originally this taxon, isolated from fruit <strong>and</strong> sports drinks,<br />

was thought to be an undescribed species of Pseudocladosporium<br />

(= Fusicladium, see below). However, upon closer examination, this<br />

197


Crous et al.<br />

Fig. 11. Cladophialophora proteae (<strong>CBS</strong> 111667). A. Colony on OA. B–C. Conidiophores. D–H. Catenulate conidia. Scale bars = 10 µm.<br />

proved not to be the case. Conidiophores appear as distinct tufts<br />

under the dissecting microscope, <strong>and</strong> are readily distinguishable<br />

from the superficial mycelium, as is normally observed in species<br />

of Fusicladium, but the conidial chains are extremely long, <strong>and</strong> the<br />

conidia tend to be more ellipsoid than the predominantly fusiform or<br />

subcylindrical conidia observed in species of Fusicladium. Hyphal<br />

coils were also not observed in cultures of C. potulentorum, but<br />

are rather common in species of Fusicladium. <strong>The</strong> phylogenetic<br />

position of this taxon within the Herpotrichiellaceae clade also<br />

supports inclusion in the <strong>genus</strong> Cladophialophora.<br />

Cladophialophora proteae Viljoen & Crous, S. African J. Bot. 64:<br />

137. 1998. Fig. 11.<br />

≡ Pseudocladosporium proteae (Viljoen & Crous) Crous, in Crous et al.,<br />

Cultivation <strong>and</strong> Diseases of Proteaceae: Leucadendron, Leucospermum<br />

<strong>and</strong> Protea: 101. 2004.<br />

In vitro: Mycelium consisting of branched, septate hyphae, often<br />

forming str<strong>and</strong>s, anastomosing, smooth to finely verruculose,<br />

frequently constricted at septa, olivaceous, 3–4 µm wide; hyphal<br />

cells in older cultures becoming swollen, up to 6 µm wide.<br />

Conidiophores reduced to conidiogenous cells. Conidiogenous<br />

cells holoblastic, integrated, forming short, truncate protuberances,<br />

2–3 × 1.5–2 µm, concolorous with mycelium, subcylindrical.<br />

Conidia in vitro arranged in long acropetal chains (up to 20), simple<br />

or branched, subcylindrical to oblong-doliiform, (9–)13–17(–22) ×<br />

2.5–3(–4) µm in vitro on MEA, (9–)16–22(–25) × (2.5–)3–4(–6)<br />

µm on SNA; 0–1(–4)-septate, pale brown to pale olivaceous,<br />

smooth, hila subtruncate to truncate, not thickened, but somewhat<br />

refractive.<br />

Cultural characteristics: Colonies erumpent, with sparse aerial<br />

mycelium on PDA; margins irregular, feathery; greyish rose, with<br />

patches of pale olivaceous-grey (surface); reverse olivaceous-grey.<br />

Colonies reaching 10 mm diam after 2 wk at 25 °C in the dark;<br />

colonies fertile.<br />

Specimen examined: South Africa, Western Cape Province, Stellenbosch, J.S.<br />

Marais Nature Reserve, leaves of Protea cynaroides (Proteaceae), 26 Aug. 1996, L.<br />

Viljoen, holotype PREM 55345, culture ex-type <strong>CBS</strong> 111667.<br />

Notes: Cladophialophora proteae differs from species of Fusicladium<br />

(= Pseudocladosporium) based on its colony colour, the slimy<br />

nature of colonies, as well as its conidia that have inconspicuous,<br />

unthickened hila (Fig. 11) (Crous et al. 2004), unlike those observed<br />

in species of Fusicladium. Sequence data show that this species is<br />

not allied to the Venturiaceae, but to the Herpotrichiellaceae.<br />

Cladophialophora scillae (Deighton) Crous, U. Braun & K. Schub.,<br />

comb. nov. MycoBank MB504530. Fig. 12.<br />

Basionym: <strong>Cladosporium</strong> scillae Deighton, N. Zeal<strong>and</strong> J. Bot. 8:<br />

55. 1970.<br />

≡ Fusicladium scillae (Deighton) U. Braun & K. Schub., IMI Descriptions of<br />

Fungi <strong>and</strong> Bacteria 152: 1518. 2002.<br />

198


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

Fig. 12. Cladophialophora scillae (<strong>CBS</strong> 116461). A–C. Conidiophores. D–F. Catenulate conidia. Scale bar = 10 µm.<br />

In vivo: see Schubert & Braun (2002a) <strong>and</strong> Schubert et al. (2003).<br />

In vitro: Mycelium consisting of branched, septate, smooth, greenbrown<br />

to medium brown, guttulate hyphae, variable in width, 1.5–3<br />

µm diam. Conidiophores lateral or terminal on hyphae, erect,<br />

straight to slightly flexuous, solitary, in some cases aggregated,<br />

subcylindrical, curved to geniculate-sinuous, unbranched, up to 55<br />

µm long, 2–3 µm wide, 0–7-septate, septa in short succession,<br />

pale to medium brown, somewhat paler towards apices, smooth.<br />

Conidiogenous cells integrated, terminal or lateral as individual loci<br />

on hyphal cells, straight to curved, subcylindrical, up to 14(–18) µm<br />

long <strong>and</strong> 2 µm wide, pale to medium brown, smooth, with a single or<br />

few subdenticulate to denticulate loci at the apex due to sympodial<br />

proliferation, or reduced to individual loci, 0.8–1.5(–2) µm wide;<br />

scars minutely thickened <strong>and</strong> darkened, but not refractive. Conidia<br />

occurring in long, unbranched or loosely branched chains (–30),<br />

straight to slightly curved, ellipsoid to mostly narrowly subcylindrical,<br />

obclavate in some larger, septate conidia, (5–)10–20(–35) × 1.5–3<br />

µm, 0–1(–3)-septate, sometimes slightly constricted at the septa,<br />

subhyaline to pale brown, smooth, guttulate, tapering at ends<br />

to subtruncate hila, 0.8–1.5 µm wide, minutely thickened <strong>and</strong><br />

darkened, but not refractive; microcyclic conidiogenesis occurring.<br />

Cultural characteristics: Colonies erumpent, spreading, with<br />

smooth, even margins <strong>and</strong> dense, abundant aerial mycelium on<br />

PDA; grey-olivaceous (surface); reverse dark olivaceous. Colonies<br />

on OA olivaceous-grey, smoke-grey due to profuse sporulation,<br />

www.studiesinmycology.org<br />

reverse olivaceous-grey to iron-grey, velvety, aerial mycelium<br />

sparse, diffuse. Colonies reaching 20 mm diam on SNA, <strong>and</strong> 40<br />

mm on PDA after 1 mo at 25 °C in the dark; colonies fertile.<br />

Specimens examined: New Zeal<strong>and</strong>, Levin, on Scilla peruviana (Hyacinthaceae),<br />

21 Dec. 1965, G.F. Laudon, IMI 116997 holotype; Auckl<strong>and</strong>, Manurewa, Auckl<strong>and</strong><br />

Botanic Gardens, on leaf spots of Scilla peruviana, 25 Apr. 2004, C.F. Hill, 1044,<br />

<strong>CBS</strong> H-19903, epitype designated here, culture ex-type <strong>CBS</strong> 116461.<br />

Notes: In culture Cladophialophora scillae forms a<br />

pseudocladosporium-like state, though the scars are somewhat<br />

darkened <strong>and</strong> thickened, but not refractive. Conidiophores are<br />

reduced to conidiogenous cells that are integrated in the mycelium,<br />

terminal or lateral, frequently also as an inconspicuous lateral<br />

denticle, with a flat-tipped scar. Conidia occur in long, branched<br />

chains, which are subcylindrical to narrowly ellipsoid, <strong>and</strong> are<br />

up to 35 µm long, 1.5–3 µm wide, thus longer <strong>and</strong> thinner than<br />

reported on the host, which were 0–3-septate, subcylindrical to<br />

ellipsoid-ovoid, 7–22 × 2.5–4 µm. Due to the fusicladioid habit of<br />

this species in vivo, Schubert & Braun (2002a) reallocated it to<br />

Fusicladium. Based on ITS sequence data, morphology <strong>and</strong> cultural<br />

characteristics, Cladophialophora scillae was almost identical to<br />

an isolate obtained from leaf spots of Hosta plantaginea in Korea.<br />

<strong>The</strong>se isolates appeared to resemble species of Fusicladium, but<br />

phylogenetically they clustered in the Herpotrichiellaceae. <strong>The</strong>refore,<br />

“Fusicladium” scillae was placed in the <strong>genus</strong> Cladophialophora.<br />

As far as we are aware, this species <strong>and</strong> C. hostae are first reports<br />

of phytopathogenic species within the <strong>genus</strong> Cladophialophora.<br />

199


Crous et al.<br />

Fig. 13. Cladophialophora sylvestris (<strong>CBS</strong> 350.83). A–B. Conidiophores. C. Catenulate conidia. D. Conidial mass. Scale bar = 10 µm.<br />

Fig. 14. Cyphellophora hylomeconis (<strong>CBS</strong> 113311). A. Colony on PDA. B–C. Hyphae with truncate conidiogenous loci. D–F. Conidia. Scale bars = 10 µm.<br />

Cladophialophora sylvestris Crous & de Hoog, sp. nov.<br />

MycoBank MB504531. Fig. 13.<br />

Etymology: Refers to its host, Pinus sylvestris.<br />

Cladophialophorae humicolae similis, sed conidiis 0–3-septatis, (7–)10–16(–20) ×<br />

1.5–2 µm.<br />

Mycelium composed of branched, smooth, pale olivaceous to<br />

pale brown hyphae, frequently forming hyphal coils, not to slightly<br />

constricted at the septa, 1–2 µm wide. Conidiophores medium<br />

brown, subcylindrical, flexuous, mononematous, multiseptate, up<br />

to 50 µm long, <strong>and</strong> 2–3 µm wide. Conidiogenous cells apical,<br />

sympodial, pale brown, 5–12 × 2–3 µm; scars somewhat darkened<br />

<strong>and</strong> thickened, not refractive. Conidia occurring in branched chains;<br />

ramoconidia up to 2 µm wide, giving rise apically to disarticulating<br />

chains of conidia; smooth, 0–3-septate, pale olivaceous,<br />

subcylindrical, (7–)10–16(–20) × 1.5–2 µm, with truncate ends; hila<br />

somewhat darkened <strong>and</strong> thickened, not refractive.<br />

Cultural characteristics: Colonies erumpent on PDA, with smooth,<br />

catenulate margins; iron-grey (surface); reverse greenish black.<br />

Colonies reaching 15 mm diam after 1 mo at 25 °C in the dark;<br />

colonies fertile.<br />

Specimen examined: Netherl<strong>and</strong>s, Kootwijk, needle litter of Pinus sylvestris<br />

(Pinaceae), 8 Nov. 1982, G.S. de Hoog, holotype <strong>CBS</strong> H-19917, culture ex-type<br />

<strong>CBS</strong> 350.83.<br />

Notes: Morphologically <strong>CBS</strong> 350.83 was originally identified as<br />

Polyscytalum griseum Sacc., but the latter is reported to have<br />

conidia that are 5–5.5 × 1 µm (Saccardo 1877), which is much<br />

smaller than that observed for the present isolate. Furthermore,<br />

the type species of Polyscytalum, P. fecundissimum Riess (<strong>CBS</strong><br />

100506), does not cluster within the Herpotrichiellaceae, thus<br />

suggesting that <strong>CBS</strong> 350.83 is best treated as a new species of<br />

Cladophialophora.<br />

Cyphellophora hylomeconis Crous, de Hoog & H.D. Shin, sp.<br />

nov. MycoBank MB504532. Fig. 14.<br />

Etymology: Named after its host <strong>genus</strong>, Hylomecon.<br />

Cyphellophorae lacinatae similis, sed conidiis longioribus et leniter angustioribus,<br />

(15–)25–35(–55) × (2.5–)3(–4) µm.<br />

Mycelium consisting of branched, greenish brown, septate,<br />

branched, smooth, 3–5 µm wide hyphae, constricted at septa.<br />

Conidiogenous cells phialidic, intercalary, appearing denticulate,<br />

1 µm tall, 1.5–2 µm wide, with minute collarettes (at times<br />

200


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

proliferating percurrently). Conidia sickle-shaped, smooth, medium<br />

brown, guttulate, (1–)3(–5)-septate, constricted at septa, widest in<br />

middle, or lower third of the conidium; apex subacutely rounded,<br />

base subtruncate, or having a slight constriction, giving rise to a<br />

foot cell, 1 µm long, 0.5–1 µm wide, subacutely rounded, (15–)25–<br />

35(–55) × (2.5–)3(–4) µm; a marginal frill is visible above the foot<br />

cell, suggesting this foot cell may be the onset of basal germination;<br />

conidia also anastomose <strong>and</strong> undergo microcyclic conidiation in<br />

culture.<br />

Cultural characteristics: Colonies slow-growing, slimy, aerial<br />

mycelium absent, margins smooth, catenate; surface crumpled,<br />

olivaceous-black to iron-grey. Colonies reaching 20 mm diam<br />

after 1 mo at 25 °C in the dark on PDA, 12 mm on SNA; colonies<br />

fertile.<br />

Specimen examined: Korea, Yangpyeong, on leaves of Hylomecon verlance<br />

(Papaveraceae), 4 Jun. 2003, H.D. Shin, holotype <strong>CBS</strong> H-19907, isotype SMK<br />

19550, culture ex-type <strong>CBS</strong> 113311.<br />

Notes: Cyphellophora hylomeconis is related to the type species<br />

of the <strong>genus</strong>, Cyphellophora laciniata G.A. de Vries, which also<br />

resides in the Herpotrichiellaceae. <strong>The</strong> <strong>genus</strong> Cyphellophora G.A.<br />

de Vries is phenetically distinguished from Pseudomicrodochium<br />

B.C. Sutton, typified by P. aciculare B.C. Sutton (1975) by melanized<br />

versus hyaline thalli. Phylogenetic confirmation is pending due to<br />

unavailability of sequence data. Decock et al. (2003) synonymised<br />

the hyaline <strong>genus</strong> Kumbhamaya M. Jacob & D.J. Bhat (Jacob &<br />

Bhat 2000) with Cyphellophora, but as no cultures of this fungus are<br />

available this decision seems premature. Nearly all Cyphellophora<br />

species accepted by Decock et al. (2003) have been found to be<br />

involved in cutaneous infections in humans. This also holds true<br />

for the species originally described as being environmental, C.<br />

vermispora Walz & de Hoog, which is closely related to C. suttonii<br />

(Ajello et al.) Decock <strong>and</strong> C. fusarioides (C.K. Campbell & B.C.<br />

Sutton) Decock known from proven human <strong>and</strong> animal infections.<br />

Decock et al. (2003) added the melanized species C. guyanensis<br />

Decock & Delgado, isolated as a saprobe from tropical leaf litter.<br />

Cyphellophora hylomeconis is the first species of the <strong>genus</strong><br />

infecting a living plant host. ITS sequences are remote from those of<br />

the remaining Cyphellophora species, the nearest neighbour being<br />

C. pluriseptata G.A. de Vries, Elders & Luykx at 19.1 % distance<br />

(data not shown). Cyphellophora hylomeconis can be distinguished<br />

based on its conidial dimensions <strong>and</strong> septation. Conidia are larger<br />

than those of C. fusarioides (11–20 × 2–2.5 µm, 1–2-septate), <strong>and</strong><br />

those of C. laciniata (11–25 × 2–5 µm, 1–3-septate) (for a key to<br />

the species see Decock et al. 2003).<br />

Exophiala sp. 1. Fig. 15.<br />

Mycelium consisting of smooth, branched, septate, medium brown,<br />

2–3 µm wide hyphae, regular in width, forming hyphal str<strong>and</strong>s<br />

<strong>and</strong> hyphal coils, with free yeast-like cells present in culture;<br />

chlamydospores terminal on hyphae, frequently forming clusters<br />

or chains, medium brown, ellipsoid, 0–1-septate, up to 10 µm<br />

long <strong>and</strong> 5 µm wide. Conidiophores reduced to conidiogenous<br />

cells, or consisting of one supporting cell, giving rise to a single<br />

conidiogenous cell, subcylindrical to ellipsoid, medium brown,<br />

smooth, 5–12 × 3.5–4 µm, with 1(–3) phialidic loci, somewhat<br />

protruding, appearing subdenticulate at first glance under the<br />

light microscope. Conidiogenous cells integrated as lateral loci on<br />

hyphal cells, inconspicuous, 1–1.5 µm wide, with a slightly flaring<br />

collarette, (1–)1.5(–2) µm long. Conidia ellipsoid, smooth, guttulate,<br />

becoming brown, swollen <strong>and</strong> elongated, <strong>and</strong> at times 1-septate,<br />

4–5(–7) × (2.5–)3(–4) µm (description based on <strong>CBS</strong> 115142).<br />

www.studiesinmycology.org<br />

Cultural characteristics: Colonies erumpent, spreading, with sparse<br />

to dense aerial mycelium on PDA, olivaceous-grey (surface), with a<br />

thin to wide, smooth, olivaceous-black margins; reverse olivaceousblack;<br />

on OA olivaceous-grey (surface) with wide, olivaceous-black<br />

margins. Colonies reaching 40–50 mm diam after 1 mo at 25 °C in<br />

the dark; colonies fertile, but sporulation sparse. Not able to grow<br />

at 37 °C.<br />

Specimen examined: Australia, from a fruit drink, May 2002, N.J. Charley, <strong>CBS</strong><br />

115142 = CPC 11044 = FRR 5582.<br />

Notes: Species of Exophiala are frequently observed as agents<br />

of human mycoses in immunocompromised patients (de Hoog<br />

et al. 2000). <strong>The</strong>y are found in the environment as slow-growing,<br />

oligotrophic colonisers of moist substrates. For example the<br />

thermotolerant species E. dermatitidis (Kano) de Hoog <strong>and</strong> E.<br />

phaeomuriformis (Matsumoto et al.) Matos et al. are common<br />

in public steam baths (Matos et al. 2003), while E. mesophila<br />

Listemann & Freiesleben can be found in showers <strong>and</strong> swimming<br />

pools (unpubl. data). Both species are able to cause infections<br />

in humans (Zeng et al. 2007). Several other species have been<br />

associated primarily with infections in fish <strong>and</strong> cold-blooded animals<br />

(Richards et al. 1978) <strong>and</strong> are occasionally found on humans<br />

(Madan et al. 2006). <strong>The</strong> occurrence of the present species in fruit<br />

drinks, therefore, is cause of concern, although it was unable to<br />

grow at 37 °C. This species forms part of a larger study, <strong>and</strong> will be<br />

treated elsewhere.<br />

Exophiala sp. 2. Fig. 16.<br />

Mycelium consisting of smooth, branched, septate, pale brown,<br />

1.5–3 µm wide hyphae, forming hyphal str<strong>and</strong>s <strong>and</strong> hyphal coils;<br />

hyphae at times terminating in chains of ellipsoid chlamydospores<br />

that are medium brown, smooth, up to 10 µm long <strong>and</strong> 5 µm wide.<br />

Conidiophores subcylindrical, medium brown, smooth, consisting<br />

of a supporting cell <strong>and</strong> a single conidiogenous cell, or reduced<br />

to a conidiogenous cell, straight to curved, up to 30 µm long<br />

<strong>and</strong> 2–3 µm wide. Conidiogenous cells pale to medium brown,<br />

subcylindrical to narrowly ellipsoid or subclavate, with 1–3 apical,<br />

phialidic loci, 1 µm wide, 1–2 µm tall, collarette somewhat flaring,<br />

but mostly cylindrical, 7–20 × 2–2.5 µm; at times proliferating<br />

percurrently. Conidia ellipsoid, smooth, guttulate, hyaline, becoming<br />

pale olivaceous, apex obtuse, base subtruncate, (4–)5–7(–10) ×<br />

2–2.5(–3) µm.<br />

Cultural characteristics: Colonies spreading with smooth,<br />

submerged margins, moderate aerial mycelium on PDA, sparse on<br />

OA, on PDA <strong>and</strong> OA olivaceous-grey (surface), with a wide, irongrey<br />

margin; reverse iron-grey. Colonies reaching 40–50 mm diam<br />

after 1 mo at 25 °C in the dark; colonies fertile. Not able to grow<br />

at 37 °C.<br />

Specimen examined: Australia, from bottled spring water, May 2003, N.J. Charley,<br />

<strong>CBS</strong> 115143 = CPC 11047 = FRR 5599.<br />

Notes: This strain represents another taxon occurring in bottled<br />

drinks destined for human consumption. As it is unable to grow at<br />

37 °C, it does not appear to pose any serious threat to human<br />

health. This species forms part of a larger study, <strong>and</strong> will be treated<br />

elsewhere.<br />

201


Crous et al.<br />

Fig. 15. Exophiala sp. 1 (<strong>CBS</strong> 115142). A. Colony on PDA. B. Hyphal coil. C. Hyphal str<strong>and</strong>. D–H. Conidiogenous cells <strong>and</strong> loci. I–O. Conidiogenous cells <strong>and</strong> conidia. Scale<br />

bars = 10 µm.<br />

Fig. 16. Exophiala sp. 2 (<strong>CBS</strong> 115143). A. Conidiogenous cells. B. Conidiophore with hyphal coil. C. Conidiogenous cell with hyphal str<strong>and</strong>. D. Conidia. Scale bar = 10 µm.<br />

202


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

Fig. 17. Exophiala eucalyptorum (CPC 11261). A. Colony on PDA. B–H. Hyphae, conidiogenous cells <strong>and</strong> conidia. Scale bars = 10 µm.<br />

Exophiala eucalyptorum Crous, sp. nov. MycoBank MB504533.<br />

Fig. 17.<br />

Etymology: Named after its occurrence on Eucalyptus leaves.<br />

Exophialae spiniferae similis, sed conidiis fusoidibus-ellipsoideis, (5–)6–8(–10)<br />

× (3–)4–5(–7) µm, et cellulis conidiogenis saepe catenatis, in catenis brevibus,<br />

dividentibus.<br />

Mycelium consisting of smooth to finely verruculose, branched,<br />

septate, 2–4 µm wide hyphae, at times giving rise to chains of<br />

dark brown, fusoid-ellipsoid chlamydospores, which can still have<br />

phialides, suggesting they were conidiogenous cells; hyphae<br />

becoming constricted at septa when fertile. Conidiophores reduced<br />

to conidiogenous cells. Conidiogenous cells numerous, terminal<br />

<strong>and</strong> lateral, mono- to polyphialidic, 5–15 × 3–5 µm; loci 1–1.5 µm<br />

wide <strong>and</strong> tall, with inconspicuous collarettes, at time proliferating<br />

percurrently; conidiogenous cells fusoid-ellipsoid, <strong>and</strong> frequently<br />

breaking off, appearing as short chains of conidia, but distinct in<br />

having conidiogenous loci. Conidia fusoid-ellipsoid, apex acutely<br />

rounded, base subtruncate, (5–)6–8(–10) × (3–)4–5(–7) µm;<br />

frequently becoming fertile, septate <strong>and</strong> brown with age.<br />

Cultural characteristics: Colonies erumpent, convex, smooth, slimy,<br />

margins feathery to crenate <strong>and</strong> smooth; aerial mycelium absent,<br />

growth yeast-like. Colonies on PDA, OA <strong>and</strong> SNA chestnut on<br />

surface <strong>and</strong> reverse. Colonies reaching 4 mm diam after 2 wk on<br />

PDA at 25 °C in the dark.<br />

Specimen examined: New Zeal<strong>and</strong>, Wellington Botanical Garden, on leaf litter of<br />

Eucalyptus sp. (Myrtaceae), Mar. 2004, J.A. Stalpers, holotype <strong>CBS</strong> H-19905,<br />

culture ex-type <strong>CBS</strong> 121638 = CPC 11261.<br />

www.studiesinmycology.org<br />

Notes: Exophiala eucalyptorum is rather characteristic in that,<br />

in culture, chains of conidiogenous cells frequently detach from<br />

hyphae, appearing as short, intact chains of fertile conidia.<br />

Its phylogenetic position is somewhat outside the core of the<br />

Herpotrichiellaceae containing most Capronia teleomorphs <strong>and</strong><br />

the remaining opportunistic Exophiala species, but still within the<br />

Chaetothyriales (Figs 1–2).<br />

Members of Venturiaceae<br />

Anungitea B. Sutton <strong>and</strong> Anungitopsis R.F. Castañeda &<br />

W.B. Kendr.<br />

Sutton (1973) erected the <strong>genus</strong> Anungitea to accommodate<br />

species with brown, mononematous conidiophores bearing<br />

apically aggregated, flat-tipped, subdenticulate conidiogenous loci<br />

that give rise to chains of pale brown subcylindrical conidia with<br />

thickened, darkened hila. He compared the type species, A. fragilis<br />

B. Sutton with anamorph genera of the Mycosphaerellaceae,<br />

but did not compare it to Fusicladium, to which it is remarkably<br />

<strong>similar</strong>. Castañeda & Kendrick (1990b) introduced the <strong>genus</strong><br />

Anungitopsis based on A. speciosa R.F. Castañeda & W.B. Kendr.<br />

This <strong>genus</strong> was distinguished from Anungitea by its formation of<br />

subdenticulate conidiogenous loci distributed along the apical<br />

region of the conidiophore, <strong>and</strong> by the relatively poorly defined<br />

appearance of these loci. No cultures are available of the extype<br />

species of Anungitea, but we studied strains of Anungitopsis<br />

203


Crous et al.<br />

Fig. 18. Cylindrosympodium lauri (<strong>CBS</strong> 240.95). A–C. Conidiophores with conidiogenous loci. D. Conidia. Scale bar = 10 µm.<br />

amoena R.F. Castañeda & Dugan (<strong>CBS</strong> 254.95, ex-type), <strong>and</strong><br />

Anungitopsis intermedia Crous & W.B. Kendr. (<strong>CBS</strong> 110746,<br />

ex-epitype), <strong>and</strong> found them to cluster adjacent to Fusicladium<br />

(Venturiaceae). However, the ex-type strain of Anungitopsis<br />

speciosa (<strong>CBS</strong> 181.95), type species of Anungitopsis, clustered<br />

distantly from all other species, confirming that the <strong>genus</strong> name<br />

Anungitopsis is not available for any of the taxa treated here. In<br />

any case, A. speciosa has unusual subdenticulate conidiogenous<br />

loci with indistinct marginal frills, <strong>and</strong> these are obviously different<br />

from those of anungitea- <strong>and</strong> fusicladium-like anamorphs, including<br />

A. amoena <strong>and</strong> A. intermedia. <strong>The</strong> latter two species previously<br />

referred to as Anungitopsis belong to a sister clade of the Venturia<br />

(Fusicladium, incl. Pseudocladosporium) clade. Sympoventuria<br />

(Crous et al. 2007b), which produces a sympodiella-like anamorph<br />

in culture, is the only teleomorph of this clade hitherto known. <strong>The</strong><br />

venturia-like habit of Sympoventuria, connected with fusicladium-<br />

/ pseudocladosporium-like anamorphs distributed in both clades,<br />

indicates a close relation between these clades, suggesting a<br />

placement in the Venturiaceae. Schubert et al. (2003) referred to<br />

the difficulty to distinguish between Anungitea <strong>and</strong> Fusicladium.<br />

Anungitea is undoubtedly heterogeneous. Anungitea rhabdospora<br />

P.M. Kirk (Kirk 1983) is, for instance, intermediate between<br />

Anungitea (conidiophores with a terminal denticulate conidiogenous<br />

cell, but conidia disarticulating in an arthroconidium-like manner)<br />

<strong>and</strong> Sympodiella B. Kendr. (conidiophores distinctly sympodial,<br />

forming arthroconidia). Other species assigned to Anungitea<br />

possess a distinctly swollen, lobed conidiophore base, e.g. A.<br />

heterospora P.M. Kirk (Kirk 1983), which is comparable with other<br />

morphologically <strong>similar</strong> genera, e.g., Parapleurotheciopsis P.M.<br />

Kirk (Kirk 1982), Rhizocladosporium Crous & U. Braun (see Crous<br />

et al. 2007a – this volume), <strong>and</strong> Subramaniomyces Varghese &<br />

V.G. Rao (Varghese & Rao 1979, Kirk 1982). <strong>The</strong> application of<br />

Anungitea depends, however, on the affinity of A. fragilis, the type<br />

species, of which sequence data are not yet available. <strong>The</strong> best<br />

solution for this problem is the widened application of Fusicladium<br />

(incl. Pseudocladosporium) to both sister clades, i.e., to the whole<br />

Venturiaceae. Morphologically a distinction between fusicladioid<br />

anamorphs of both clades is impossible. <strong>The</strong> more “fusicladium-like”<br />

growth is mainly characteristic for the fruiting in vivo, above all in<br />

biotrophic taxa, whereas the more “pseudocladosporium-like” habit<br />

is typical for the growth in vitro <strong>and</strong> in saprobic taxa, a phenomenon<br />

which is also evident in species of the morphologically <strong>similar</strong><br />

<strong>genus</strong> Cladophialophora (see C. hostae <strong>and</strong> C. scillae). A potential<br />

placement of Anungitea fragilis within the Venturiaceae, which has<br />

still to be proven, would render the <strong>genus</strong> Anungitea a synonym of<br />

Fusicladium, but in the case of a quite distinct phylogenetic position<br />

a new circumscription of this <strong>genus</strong>, excluding the Venturiaceae<br />

anamorphs, would be necessary. Thus, a final conclusion about<br />

Anungitea has to be postponed, awaiting cultures <strong>and</strong> sequence<br />

analyses of its type species.<br />

<strong>The</strong> taxonomic placement of a fungus from the Canary<br />

Isl<strong>and</strong>s, isolated from leaf litter of Laurus sp. (<strong>CBS</strong> 240.95), is<br />

somewhat problematic. It clusters within the Venturiaceae, but<br />

not within Venturia s. str. itself, <strong>and</strong> it does not fit into the current<br />

morphological concept of Fusicladium (incl. Pseudocladosporium).<br />

Based on its solitary, cylindrical, hyaline conidia <strong>and</strong> pale brown<br />

conidiogenous structures, it resembles species accommodated in<br />

Cylindrosympodium W.B. Kendr. & R.F. Castañeda (Castañeda &<br />

Kendrick 1990a, Marvanová & Laichmanová 2007).<br />

Cylindrosympodium lauri Crous & R.F. Castañeda, sp. nov.<br />

MycoBank MB504534. Fig. 18.<br />

Etymology: Named after the host <strong>genus</strong> it was collected from,<br />

Laurus.<br />

Cylindrosympodii variabilis similis, sed conidiophoris longioribus, ad 70 µm, conidiis<br />

subhyalinis vel dilute olivaceis.<br />

Mycelium consisting of brown, smooth, septate, branched hyphae,<br />

1.5–2.5 µm wide. Conidiophores macronematous, mononematous,<br />

solitary, erect, subcylindrical, straight to geniculate-sinuous,<br />

medium brown, smooth, 35–70 × 2.5–4 µm, 1–5-septate.<br />

Conidiogenous cells terminal, integrated, pale to medium brown,<br />

smooth, 10–35 × 2–3 µm, proliferating sympodially, with one to<br />

several flat-tipped loci, 1.5–2 µm wide; scars somewhat darkened,<br />

minutely thickened, but not refractive. Conidia solitary, subacicular<br />

to narrowly subcylindrical, apex subobtuse, base truncate, or<br />

somewhat swollen, straight or curved, smooth, subhyaline to very<br />

pale olivaceous, guttulate, (45–)60–70(–80) × 2.5–3(–3.5) µm,<br />

(4–)6–8-septate; scars are somewhat darkened, minutely thickened,<br />

but not refractive, 2.5–3 µm wide.<br />

Cultural characteristics: Colonies erumpent, convex, with smooth,<br />

lobed margins, <strong>and</strong> moderate, dense aerial mycelium on PDA;<br />

mouse-grey in the central part, <strong>and</strong> dark mouse-grey in the outer<br />

zone (surface); reverse dark mouse-grey. Colonies reaching 5 mm<br />

diam after 2 wk at 25 °C in the dark; colonies fertile.<br />

204


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

Specimen examined: Spain, Canary Isl<strong>and</strong>s, leaf litter of Laurus sp. (Lauraceae), 4<br />

Jan. 1995, R.F. Castañeda, holotype <strong>CBS</strong> H-19909, culture ex-type <strong>CBS</strong> 240.95.<br />

Note: <strong>The</strong> present fungus differs from Cylindrosympodium variabile<br />

(de Hoog) W.B. Kendr. & R.F. Castañeda (de Hoog 1985) in that the<br />

conidiophores are much longer, the conidia are subhyaline to very<br />

pale olivaceous, <strong>and</strong> the scars <strong>and</strong> hila are thin, slightly darkened,<br />

but not refractive.<br />

Venturia Sacc. <strong>and</strong> its anamorph Fusicladium<br />

Venturia Sacc., Syll. fung. (Abellini) 1: 586. 1882.<br />

= Apiosporina Höhn., Sitzungsber. Kaiserl. Akad. Wiss., Math.-Naturw. Cl., Abt.<br />

1, 119: 439. 1910, syn. nov.<br />

= Metacoleroa Petr., Ann. Mycol. 25: 332. 1927, syn. nov.<br />

= Caproventuria U. Braun, A Monograph of Cercosporella, Ramularia <strong>and</strong> Allied<br />

Genera (Phytopathogenic Hyphomycetes) 2: 396. 1998, syn. nov.<br />

For additional synonyms see Sivanesan, <strong>The</strong> bitunicate<br />

Ascomycetes <strong>and</strong> their anamorphs: 604. 1984.<br />

Anamorph: Fusicladium Bonord., H<strong>and</strong>b. Mykol.: 80. 1851.<br />

= Pseudocladosporium U. Braun, A Monograph of Cercosporella, Ramularia<br />

<strong>and</strong> Allied Genera (Phytopathogenic Hyphomycetes) 2: 392. 1998, syn. nov.<br />

For additional synonyms, see Schubert et al. (2003).<br />

Notes: <strong>The</strong> <strong>genus</strong> Caproventuria, based on C. hanliniana (U. Braun<br />

& Feiler) U. Braun, was erected to accommodate saprobic, soil-borne<br />

venturia-like ascomycetes with numerous ascomatal setae, <strong>and</strong> an<br />

anamorph quite distinct from Fusicladium (Braun 1998). <strong>The</strong> <strong>genus</strong><br />

Metacoleroa is based on M. dickiei (Berk. & Broome) Petr., which<br />

clusters in the Venturiaceae, adjacent to Caproventuria, which has<br />

Pseudocladosporium anamorphs. Metacoleroa was retained by Barr<br />

(1987) as separate from Venturia based on its superficial ascomata<br />

with a thin, stromatic layer beneath the ascomata. Whether these<br />

criteria still justify the separation of Caproventuria <strong>and</strong> Metacoleroa<br />

from Venturia is debatable, <strong>and</strong> the names Venturia dickiei (Berk. &<br />

Broome) Ces. & de Not. <strong>and</strong> Venturia hanliniana (U. Braun & Feiler)<br />

Unter. are available for these organisms. <strong>The</strong> <strong>genus</strong> Apiosporina,<br />

which is based on Apiosporina collinsii (Schwein.) Höhn., clusters in<br />

the Venturiaceae, as was to be expected based on its Fusicladium<br />

anamorph (Schubert et al. 2003). It was distinguished from Venturia<br />

species by having ascospores strictly septate near the lower end<br />

(Sivanesan 1984).<br />

<strong>The</strong> anamorph <strong>genus</strong> Fusicladium has been monographed by<br />

Schubert et al. (2003). Morphological as well as molecular studies<br />

(Beck et al. 2005) demonstrated that the <strong>genus</strong> Venturia with its<br />

Fusicladium anamorphs is monophyletic. A separation of Venturia<br />

into various uniform subclades based on the previous anamorph<br />

genera Fusicladium, Pollaccia <strong>and</strong> Spilocaea was not evident <strong>and</strong><br />

could be rejected. As in cercosporoid anamorphs of Mycosphaerella,<br />

features such as the arrangement of the conidiophores (solitary,<br />

fasciculate, sporodochial), the proliferation of conidiogenous cells<br />

(sympodial, percurrent) <strong>and</strong> shape, size as well as formation of<br />

conidia (solitary, catenate) proved to be of little taxonomic value at<br />

generic level. Hence, Schubert et al. (2003) proposed to maintain<br />

Fusicladium emend. as sole anamorph <strong>genus</strong> for Venturia.<br />

<strong>The</strong> <strong>genus</strong> Fusicladosporium Partridge & Morgan-Jones (type<br />

species: <strong>Cladosporium</strong> carpophilum Thüm.) (Partridge & Morgan-<br />

Jones 2003), recently erected to accommodate fusicladium-like<br />

species with catenate conidia, represents a further synonym of<br />

Fusicladium.<br />

Similar to their occurrence in vivo the conidiophores in<br />

vitro of species previously referred to the genera Spilocaea<br />

<strong>and</strong> Pollaccia are usually micronematous, conidia often appear<br />

to be directly formed on the mycelium, unilocal, determinate,<br />

mostly reduced to conidiogenous cells, sometimes forming a few<br />

percurrent proliferations, whereas the conidiophores of species<br />

of Fusicladium s. str. are mostly macronematous, but sometimes<br />

also micronematous. <strong>The</strong>y are often initiated as short lateral,<br />

peg-like outgrowths of hyphae which proliferate sympodially,<br />

becoming slightly geniculate, forming a single, several or numerous<br />

subdenticulate to denticulate, truncate, unthickened or only slightly<br />

thickened, somewhat darkened-refractive conidiogenous loci.<br />

<strong>The</strong> <strong>genus</strong> Pseudocladosporium was described to be quite<br />

distinct from Fusicladium by being saprobic <strong>and</strong> connected with a<br />

different teleomorph, viz. Caproventuria (Braun 1998). However,<br />

since the type species of Caproventuria, C. hanliniana, with its<br />

anamorph Pseudocladosporium brevicatenatum (U. Braun &<br />

Feiler) U. Braun clusters together with numerous Venturia species,<br />

the <strong>genus</strong> Pseudocladosporium should be reduced to synonymy<br />

with Fusicladium. Morphologically there is no clear delimitation<br />

between Fusicladium <strong>and</strong> Pseudocladosporium. <strong>The</strong> typically<br />

pseudocladosporium-like habit, characterised by forming solitary<br />

conidiophores, often reduced to conidiogenous cells or even<br />

micronematous, <strong>and</strong> conidia formed in long chains, is mainly found<br />

in culture, above all in saprobic taxa. <strong>The</strong> fusicladium-like growth<br />

with well-developed macronematous conidiophores is usually more<br />

evident in vivo, above all in biotrophic taxa. <strong>The</strong>re are, however, all<br />

kinds of transitions between these two genera.<br />

Fusicladium africanum Crous, sp. nov. MycoBank MB504535.<br />

Fig. 19.<br />

Etymology: Named after the continent from which it was collected,<br />

Africa.<br />

Fusicladio brevicatenato similis, sed conidiophoris brevioribus, 5–10 µm longis,<br />

conidiis minoribus, ad 20 × 3.5 µm, 0(–1)-septatis, locis conidiogenis et hilis<br />

angustioribus, 1–1.5 µm latis.<br />

Mycelium composed of smooth, medium brown, branched,<br />

septate, 1.5–2 µm wide hyphae, frequently forming hyphal coils.<br />

Conidiophores reduced to conidiogenous cells, solitary, pale to<br />

medium brown, smooth, inconspicuous, integrated in hyphae,<br />

varying from small, truncate lateral loci on hyphal cells, 1–1.5 µm<br />

wide, to micronematous conidiogenous cells, 5–10 × 2–3 µm; monoto<br />

polyblastic, sympodial, scars inconspicuous, 1 µm wide. Conidia<br />

in long, branched chains of up to 40, subcylindrical, 0(–1)-septate,<br />

pale brown, smooth; hila truncate, 1 µm wide, unthickened, neither<br />

darkened nor refractive; ramoconidia (11–)15–17(–20) × 2–3(–3.5)<br />

µm; conidia (8–)11–17 × 2–2.5 µm.<br />

Cultural characteristics: Colonies somewhat erumpent, with<br />

moderate aerial mycelium <strong>and</strong> smooth, lobate margins on PDA,<br />

ochreous to umber (surface); reverse dark umber; on OA umber;<br />

on SNA ochreous. Colonies reaching 9 mm diam on PDA after 2 wk<br />

at 25 °C in the dark; colonies fertile.<br />

Specimen examined: South Africa, Western Cape Province, Malmesbury,<br />

Eucalyptus leaf litter, Jan. 2006, P.W. Crous, holotype <strong>CBS</strong> H-19904, cultures extype<br />

CPC 12828 = <strong>CBS</strong> 121639, CPC 12829 = <strong>CBS</strong> 121640.<br />

Notes: Fusicladium africanum is a somewhat atypical member of the<br />

<strong>genus</strong>, as its conidial hila are quite unthickened <strong>and</strong> inconspicuous.<br />

Among biotrophic, leaf-spotting Fusicladium species a wider<br />

morphological variation was found pertaining to the structure of the<br />

conidiogenous loci <strong>and</strong> conidial hila, ranging from being indistinct,<br />

unthickened <strong>and</strong> not darkened-refractive to unthickened or almost<br />

so, but somewhat darkened-refractive (Schubert et al. 2003).<br />

Fusicladium africanum was found occurring with Sympoventuria<br />

capensis Crous & Seifert on Eucalyptus leaf litter in South Africa<br />

(Crous et al. 2007b).<br />

www.studiesinmycology.org<br />

205


Crous et al.<br />

Fig. 19. Fusicladium africanum (CPC 12828). A. Colony on MEA. B. Hyphal coil. C. Branched conidial chain. D–F. Conidiophores with catenulate conidia. Scale bar = 10 µm.<br />

Fig. 20. Fusicladium amoenum (<strong>CBS</strong> 254.95). A–E. Conidiophores with conidiogenous loci. F. Conidia. Scale bar = 10 µm.<br />

206


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

Fig. 21. Fusicladium convolvularum (<strong>CBS</strong> 112706). A–B, D–I. Conidiophores with conidiogenous loci. C, J–K. Ramoconidia <strong>and</strong> conidia. Scale bars = 10 µm.<br />

Fusicladium amoenum (R.F. Castañeda & Dugan) Crous, K.<br />

Schub. & U. Braun, comb. nov. MycoBank MB504536. Fig. 20.<br />

Basionym: Anungitopsis amoena R.F. Castañeda & Dugan,<br />

Mycotaxon 72: 118. 1999.<br />

≡ <strong>Cladosporium</strong> amoenum R.F. Castañeda, in Untereiner et al., 1998,<br />

nom. nud.<br />

Specimen examined: Cuba, Santiago de Cuba, La Gran Piedra, fallen leaves of<br />

Eucalyptus sp. (Myrtaceae), 2 Nov. 1994, R.F. Castañeda, (Ho et al. 1999: 117,<br />

figs 2–3) iconotype, culture ex-type <strong>CBS</strong> 254.95 = ATCC 200947 = IMI 367525 =<br />

INIFAT C94/155 = MUCL 39143.<br />

Note: In culture F. amoenum has a typical pseudocladosporium-like<br />

morphology, though the scars are neither prominently thickened,<br />

nor refractive.<br />

www.studiesinmycology.org<br />

207


Crous et al.<br />

Fig. 22. Fusicladium fagi (<strong>CBS</strong> 621.84). A. Conidiophore with truncate conidiogenous loci. B. Hypha with conidiogenous loci. C–G. Conidial chains. Scale bars = 10 µm.<br />

Fusicladium caruanianum Sacc., Ann. Mycol. 11: 20. 1913.<br />

≡ Pseudocladosporium caruanianum (Sacc.) U. Braun, Schlechtendalia<br />

9: 114. 2003.<br />

Fusicladium convolvularum Ondřej, Česká Mycol. 25: 171. 1971.<br />

Fig. 21.<br />

In vivo: Schubert et al. (2003: 37).<br />

In vitro on SNA: Mycelium unbranched or only sparingly branched,<br />

2–3 µm wide, septate, not constricted at septa, subhyaline to pale<br />

brown, smooth, walls unthickened or almost so. Conidiophores<br />

laterally arising from hyphae, erect, straight to somewhat flexuous,<br />

sometimes geniculate, unbranched, (6–)12–75 × (2.5–)3–4.5<br />

µm, aseptate or septate, pale brown or pale medium brown,<br />

smooth, walls somewhat thickened, sometimes only as short<br />

lateral conical prolongations of hyphae, occasionally irregular in<br />

shape. Conidiogenous cells integrated, terminal or conidiophores<br />

reduced to conidiogenous cells, sometimes geniculate, 6–29 µm<br />

long, proliferation sympodial, with several denticle-like loci, broadly<br />

truncate, 1.5–2(–2.5) µm wide, unthickened, somewhat refractive<br />

or darkened. Ramoconidia occurring, 20–28 × 5 µm, 0–1-septate,<br />

somewhat darker, pale medium brown, with a broadly truncate<br />

base, 3–4 µm wide, usually with several denticle-like apical loci.<br />

Conidia catenate, formed in unbranched or loosely branched<br />

chains, straight to sometimes curved, cells sometimes irregularly<br />

swollen, fusiform, subcylindrical, sometimes obpyriform, 13–35 ×<br />

3.5–5.5(–6) µm, 0–3-septate, occasionally slightly constricted at<br />

the median septum, few very large conidia with up to five septa,<br />

up to 75 µm long, 4.5–6 µm wide, subhyaline to pale brown,<br />

smooth, walls slightly thickened, slightly attenuated towards apex<br />

<strong>and</strong> base, hila broadly truncate, 1–2 µm wide, unthickened or<br />

only slightly thickened, somewhat darkened-refractive; microcyclic<br />

conidiogenesis occurring, conidia often germinating.<br />

Cultural characteristics: Colonies on PDA spreading, somewhat<br />

erumpent, with moderate aerial mycelium <strong>and</strong> regular, but feathery<br />

margins; surface fuscous black, <strong>and</strong> reverse dark fuscous black.<br />

Colonies reaching 15 mm diam after 1 mo on PDA at 25 °C in the<br />

dark.<br />

Specimens examined: Czech Republic, Libina, okraj pole pod nadrazim (okr.<br />

Sumperk), on Convolvulus arvensis (Convolvulaceae), 7 Sep. 1970, Ondřej,<br />

holotype BRA. New Zeal<strong>and</strong>, on leaves of Convolvulus arvensis, 7 Nov. 2000,<br />

C.F. Hill, epitype designated here <strong>CBS</strong> H-19911, culture ex-epitype <strong>CBS</strong> 112706<br />

= CPC 3884 = IMI 383037.<br />

Note: Conidiophores are somewhat longer <strong>and</strong> narrower in vitro<br />

than in vivo, <strong>and</strong> ramoconidia occur (Schubert & Braun 2002b,<br />

Schubert et al. 2003).<br />

208


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

Fig. 23. Fusicladium intermedium (<strong>CBS</strong> 110746). A–G. Conidiophores with sympodial conidiogenous loci. H. Conidia. Scale bar = 10 µm.<br />

Fusicladium fagi Crous & de Hoog, sp. nov. MycoBank MB504537.<br />

Fig. 22.<br />

Etymology: Named after its host, Fagus sylvatica.<br />

Fusicladio brevicatenato similis, sed conidiis secundis minoribus, (8–)11–17(–20) ×<br />

3–3.5 µm, locis conidiogenis et hilis angustioribus, 1–1.5 µm latis.<br />

Mycelium consisting of pale to medium brown, smooth to finely<br />

verruculose, branched, 2–3 µm wide hyphae. Conidiophores<br />

integrated, terminal on hyphae, 0–1-septate, mostly reduced to<br />

conidiogenous cells, also lateral, visible as small, protruding,<br />

denticle-like loci, 10–15 × 2–3.5 µm. Conidiogenous cells<br />

subcylindrical, 5–15 × 2–3.5 µm, pale to medium brown, smooth<br />

to finely verruculose, tapering to 1–3 apical loci, 1–1.5 µm wide;<br />

scars inconspicuous. Conidia pale brown, smooth, guttulate,<br />

subcylindrical to narrowly ellipsoid, occurring in simple or branched<br />

chains, 0–1(–2)-septate, tapering towards subtruncate ends, 1.5–<br />

2.5 µm wide, aseptate conidia (8–)11–17(–20) × 3–3.5 µm, septate<br />

conidia up to 40 µm long <strong>and</strong> 4 µm wide; hila inconspicuous, i.e.<br />

neither thickened nor darkened-refractive; microcyclic conidiation<br />

common in older cultures.<br />

Cultural characteristics: Colonies erumpent, spreading, with<br />

abundant aerial mycelium on PDA, <strong>and</strong> feathery to smooth margins;<br />

isabelline to patches of fuscous-black due to the absence of aerial<br />

mycelium, which collapses with age (surface); reverse fuscousblack.<br />

Colonies reaching 50 mm diam after 1 mo at 25 °C in the<br />

dark; colonies fertile.<br />

Specimen examined: Netherl<strong>and</strong>s, Baarn, Maarschalksbosch, decaying leaves of<br />

Fagus sylvatica (Fagaceae), 1 Oct. 1984, G.S. de Hoog, holotype <strong>CBS</strong> H-10366,<br />

culture ex-type <strong>CBS</strong> 621.84 = ATCC 200937.<br />

Notes: Isolate <strong>CBS</strong> 621.84 was until recently preserved at the<br />

<strong>CBS</strong> as representative of <strong>Cladosporium</strong> nigrellum Ellis & Everh., a<br />

species known from bark of Robinia sp. in the U.S.A. Morphologically<br />

it is, however, quite distinct in having somewhat larger, <strong>and</strong> more<br />

subcylindrical to ellipsoid conidia. Conidia of C. nigrellum are<br />

fusiform to limoniform, 0–3-septate, 5–15 × 4–7 µm (Ellis 1976),<br />

possessing the typical cladosporioid scars with a central convex<br />

dome <strong>and</strong> a periclinal rim which characterise it as a true member<br />

of the <strong>genus</strong> <strong>Cladosporium</strong> Link, which has been confirmed by a<br />

re-examination of type material of C. nigrellum (on inner bark of<br />

railroad ties, U.S.A., West Virginia, Fayette Co., Nuttallburg, 20<br />

Oct. 1893, L.A. Nuttall, Flora of Fayette County No. 172, NY; also<br />

Ellis & Everh., N. Amer. Fungi 3086 <strong>and</strong> Fungi Columb. 382, BPI,<br />

NY, PH).<br />

Fusicladium intermedium (Crous & W.B. Kendr.) Crous, comb.<br />

nov. Mycobank MB504538. Fig. 23.<br />

Basionym: Anungitopsis intermedia Crous & W.B. Kendr. S. Afr. J.<br />

Bot. 63: 286. 1997.<br />

Specimens examined: South Africa, Mpumalanga, from leaf litter of Eucalyptus<br />

sp. (Myrtaceae), Oct. 1992, M.J. Wingfield, PREM 51438 holotype. Madagascar,<br />

Tamatave, Eucalyptus leaf litter, Apr. 1994, P.W. Crous, <strong>CBS</strong> H-19918, epitype<br />

designated here, culture ex-epitype CPC 778 = IMI 362702 = <strong>CBS</strong> 110746.<br />

Note: Conidiophores are dimorphic in culture, being macronematous,<br />

anungitopsis-like, <strong>and</strong> micronematous, more pseudocladosporiumlike.<br />

Fusicladium matsushimae (U. Braun & C.F. Hill) Crous, U. Braun<br />

& K. Schub., comb. nov. Mycobank MB504539.<br />

Basionym: Pseudocladosporium matsushimae U. Braun & C.F. Hill,<br />

Australas. Pl. Pathol. 33: 492. 2004.<br />

www.studiesinmycology.org<br />

209


Crous et al.<br />

Fig. 24. Fusicladium pini (<strong>CBS</strong> 463.82). A–F. Conidiogenous cells with conidiogenous loci. G. Conidia. Scale bars = 10 µm.<br />

Fusicladium m<strong>and</strong>shuricum (M. Morelet) Ritschel & U. Braun,<br />

Schlechtendalia 9: 62. 2003.<br />

Basionym: Pollaccia m<strong>and</strong>shurica M. Morelet, Ann. Soc. Sci. Nat.<br />

Archéol. Toulon Var 45(3): 218. 1993.<br />

= Pollaccia sinensis W.P. Wu & B. Sutton, in herb. (IMI).<br />

Teleomorph: Venturia m<strong>and</strong>shurica M. Morelet, Ann. Soc. Sci.<br />

Nat. Archéol. Toulon Var 45(3): 219. 1993.<br />

In vivo: Schubert et al. (2003: 62).<br />

In vitro on OA: Mycelium loosely branched, filiform to narrowly<br />

cylindrical-oblong, 1–4 µm wide, later somewhat wider, up to<br />

7 µm, septate, sometimes slightly constricted at the septa,<br />

sometimes irregular in outline due to small swellings, subhyaline<br />

to pale brown, smooth, walls unthickened, sometimes aggregating,<br />

forming compact conglomerations of slightly swollen hyphal cells.<br />

Conidiophores usually reduced to conidiogenous cells, arising<br />

terminally or laterally from hyphae, subcylindrical to cylindrical,<br />

unbranched, 9–20 × (2.5–)4–5(–6) µm, aseptate, very rarely 1-<br />

septate, very pale brown, smooth, walls unthickened, monoblastic,<br />

unilocal, determinate, later occasionally becoming percurrent,<br />

enteroblastically proliferating, forming a few (up to five) annellations,<br />

loci broadly truncate, (2–)3–5 µm wide, unthickened, not darkened.<br />

Conidia solitary, straight to curved, fusiform to obclavate, distinctly<br />

apiculate, 24–45(–57) × (6–)7–9(–10.5) µm, (1–)2–4(–5)-septate,<br />

more or less constricted at septa, sometimes up to 85 µm long<br />

with up to 7 septa, septa often somewhat darkened, second cell<br />

often bulging, pale medium to medium olivaceous-brown or brown,<br />

smooth, walls somewhat thickened, somewhat attenuated towards<br />

the base, hilum broadly truncate, (2–)3–5 µm wide, unthickened,<br />

not darkened; microcyclic conidiogenesis not observed.<br />

Cultural characteristics: Colonies on OA iron-grey to olivaceousgrey<br />

due to aerial mycelium <strong>and</strong> sporulation (surface); reverse<br />

iron-grey to black, somewhat velvety; margin glabrous, olivaceous;<br />

aerial mycelium sparsely formed, loose, diffuse; sporulating.<br />

Specimens examined: China, Liaoning, on Populus simonii × P. nigra, 17 Jun. 1992,<br />

M. Morelet, holotype PC (PFN 1466); P. simonii, 20 Apr. 1993, epitype designated<br />

here <strong>CBS</strong> H-19912, culture ex-epitype <strong>CBS</strong> 112235 = CPC 3639 = MPFN 307.<br />

Note: Conidiophores are densely fasciculate in vivo, forming<br />

sporodochial conidiomata, cylindrical to ampulliform, 5–7 × 6–7.5<br />

µm (Schubert et al. 2003).<br />

Fusicladium pini Crous & de Hoog, sp. nov. MycoBank MB504540.<br />

Fig. 24.<br />

Etymology: Named after its host, Pinus.<br />

Fusicladio africano similis, sed conidiis minoribus, (6–)10–12(–17) × 1.5–2(–2.5)<br />

µm, locis conidiogenis et hilis angustioribus, 0.5–1 µm latis.<br />

Mycelium consisting of smooth, medium brown, branched,<br />

1.5–2 µm wide hyphae, giving rise to solitary, micronematous<br />

conidiophores. Conidiophores reduced to conidiogenous cells,<br />

medium to dark brown, erect, thick-walled, smooth, subcylindrical,<br />

widest at the base, tapering to a subtruncate apex, 5–15 × 2–3<br />

µm; scars flat-tipped, somewhat darkened <strong>and</strong> thickened, one to<br />

several in the apical region, somewhat protruding, 0.5–1 µm wide.<br />

Conidia in branched or unbranched chains of up to 15, medium<br />

brown, smooth, subcylindrical, 0–1-septate, widest in the middle,<br />

tapering to subtruncate ends, straight to slightly curved, (6–)10–<br />

12(–17) × 1.5–2(–2.5) µm; hila somewhat darkened <strong>and</strong> thickened,<br />

not refractive, 0.5–1 µm wide.<br />

Cultural characteristics: Colonies erumpent, with sparse aerial<br />

mycelium <strong>and</strong> smooth margins on PDA, greyish sepia (surface);<br />

reverse fuscous-black; on OA patches of greyish sepia <strong>and</strong> fuscousblack<br />

(surface); on SNA umber (surface). Colonies reaching 15 mm<br />

diam on PDA after 1 mo at 25 °C in the dark; colonies fertile.<br />

210


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

Fig. 25. Fusicladium ramoconidii (<strong>CBS</strong> 462.82). A. Colony on OA. B. Hyphal coil. C–F. Conidiophores reduced to conidiogenous cells. G–H. Conidiophores. I. Conidia. Scale<br />

bars = 10 µm.<br />

Specimen examined: Netherl<strong>and</strong>s, Baarn, De Vuursche, needle of Pinus sylvestris<br />

(Pinaceae), 12 Apr. 1982, G.S. de Hoog, holotype <strong>CBS</strong> H-1610, culture ex-type<br />

<strong>CBS</strong> 463.82.<br />

Notes: This fungus was originally maintained in the <strong>CBS</strong> collection<br />

as Anungitea uniseptata Matsush. In culture, however, only a<br />

pseudocladosporium-like state was observed. Conidiophores are<br />

reduced to conidiogenous cells, <strong>and</strong> have several apical loci as<br />

in Fusicladium, but are not subdenticulate; scars are somewhat<br />

darkened <strong>and</strong> thickened, not refractive. Conidia of F. africanum are<br />

(8–)11–17(–20) × 2–3(–3.5) µm, thus <strong>similar</strong>, but somewhat larger<br />

than the mean conidial size range (10–12 × 1.5–2 µm) observed<br />

in F. pini. <strong>The</strong> conidiogenous loci <strong>and</strong> conidial hila of F. africanum<br />

are also somewhat larger. Although the LSU sequence of F. pini is<br />

identical to that of F. ramoconidii, the ITS sequence <strong>similar</strong>ity is 97<br />

% (572/585 nucleotides).<br />

Fusicladium ramoconidii Crous & de Hoog, sp. nov. MycoBank<br />

MB504541. Figs 25–26.<br />

Etymology: Named after the presence of its characteristic<br />

ramoconidia.<br />

Fusicladio brevicatenato similis, sed ramoconidiis minoribus, (12–)15–17(–20) ×<br />

2(–3) µm, locis conidiogenis et hilis minoribus, 0.5–1 µm diam.<br />

Mycelium consisting of branched, septate, 1.5–2 µm wide hyphae,<br />

pale brown, smooth, frequently with hyphal coils. Conidiophores<br />

integrated into hyphae, <strong>and</strong> reduced to small, lateral protruding<br />

conidiogenous cells, concolorous with hyphae, or macronematous,<br />

dark brown, erect, thick-walled, 10–40 × 3–4 µm, 0–3-septate.<br />

Conidiogenous cells terminal, integrated, subcylindrical, tapering to<br />

a rounded apex, concolorous with hyphae (as hyphal pegs), or dark<br />

www.studiesinmycology.org<br />

211


Crous et al.<br />

Fusicladium rhodense Crous & M.J. Wingf., sp. nov. MycoBank<br />

MB504542. Fig. 27.<br />

Etymology: Named after the Greek Isl<strong>and</strong>, Rhodos, where it was<br />

collected.<br />

Fusicladio africano similis, sed locis conidiogenis angustioribus, 1.5–2 µm latis, et<br />

differt a F. pini ramoconidiis formantibus.<br />

Mycelium consisting of smooth to finely roughened, medium brown,<br />

branched, septate, 1.5–3 µm wide hyphae, frequently forming<br />

hyphal coils, giving rise to solitary, micronematous conidiophores.<br />

Conidiophores reduced to conidiogenous cells that are terminal<br />

or lateral on hyphae, medium brown, smooth, subcylindrical,<br />

subdenticulate, erect, or more distinct, up to 15 µm tall, 1.5–2<br />

µm wide, mono- to polyblastic; scars flat-tipped, somewhat<br />

darkened <strong>and</strong> thickened, but not refractive. Conidia in branched<br />

or unbranched chains of up to 15, pale brown in younger conidia,<br />

becoming medium brown, smooth, subcylindrical, 0–3-septate,<br />

tapering slightly towards the subtruncate ends, straight, but at times<br />

slightly curved, (8–)12–16(–20) × (2–)2.5–3(–4) µm; ramoconidia<br />

(0–)1(–3)-septate, 12–20 × 3–4 µm; conidia (0–)1-septate, 8–17<br />

× 2–3 µm; hila somewhat darkened <strong>and</strong> thickened, not refractive,<br />

1–1.5 µm wide.<br />

Fig. 26. Fusicladium ramoconidii (<strong>CBS</strong> 462.82). Conidiogenous cells with<br />

ramoconidia <strong>and</strong> conidia. Scale bar = 10 µm.<br />

brown on mononematous conidiophores, smooth, 3–15 × 2–3(–4)<br />

µm; proliferating sympodially, loci slightly thickened, darkened <strong>and</strong><br />

refractive, 0.5–1 µm wide. Conidia occurring in branched chains,<br />

narrowly ellipsoid to subcylindrical, pale olivaceous, guttulate;<br />

ramoconidia (0–)1(–3)-septate, (12–)15–17(–20) × 2(–3) µm;<br />

conidia occurring in short chains (–15), 0–1-septate, (8–)10–12(–<br />

16) × 2(–3) µm; hila slightly thickened <strong>and</strong> darkened, not refractive,<br />

0.5–1 µm wide.<br />

Cultural characteristics: Colonies erumpent, with sparse aerial<br />

mycelium <strong>and</strong> smooth margins on PDA, hazel to fawn (surface),<br />

with a thin, submerged margin; reverse brown-vinaceous; on OA<br />

hazel to fawn (surface) with a wide, fawn, submerged margin.<br />

Colonies reaching 25 mm diam on PDA after 1 mo at 25 °C in the<br />

dark; colonies fertile.<br />

Specimen examined: Netherl<strong>and</strong>s, Baarn, De Vuursche, needle of Pinus sp.<br />

(Pinaceae), 12 Apr. 1982, G.S. de Hoog, holotype <strong>CBS</strong> H-19908, culture ex-type<br />

<strong>CBS</strong> 462.82.<br />

Notes: This strain has been deposited in the <strong>CBS</strong> collection as<br />

Pseudocladosporium hachijoense (Matsush.) U. Braun. However,<br />

its ramoconidia <strong>and</strong> conidia are smaller than those cited by<br />

Matsushima (1975) (ramoconidia up to 30 µm long, conidia 10–21<br />

× 2–4 µm). Although it clusters with F. pini in the LSU phylogeny,<br />

there are 13 bp differences in their ITS sequence data. Furthermore,<br />

F. ramoconidii has ramoconidia which are absent in F. pini, <strong>and</strong><br />

has a faster growth rate, <strong>and</strong> hazel to fawn colonies, compared<br />

to the greyish sepia colonies of F. pini. <strong>The</strong> well-developed,<br />

septate conidiophores <strong>and</strong> ramoconidia are reminiscent of F.<br />

brevicatenatum, which differs, however, by its longer <strong>and</strong> wider<br />

ramoconidia, up to 30 × 6(–7) µm, as well as larger conidiogenous<br />

loci <strong>and</strong> conidial hila, 1.5–3 µm diam.<br />

Cultural characteristics: Colonies spreading, somewhat erumpent,<br />

with moderate aerial mycelium <strong>and</strong> crenate margins on PDA,<br />

uneven, greyish sepia (surface), margins fuscous-black; reverse<br />

fuscous-black; on OA smooth, spreading, with sparse aerial<br />

mycelium <strong>and</strong> even, regular margins, greyish sepia; on SNA<br />

spreading, smooth, even margins, sparse aerial mycelium, greyish<br />

sepia (surface). Colonies reaching 9 mm diam on PDA after 2 wk at<br />

25 °C in the dark; colonies fertile.<br />

Specimen examined: Greece, Rhodos, on branches of Ceratonia siliqua (Fabaceae),<br />

1 Jun. 2006, P.W. Crous & M.J. Wingfield, holotype <strong>CBS</strong> H-19910, culture ex-type<br />

<strong>CBS</strong> 121641 = CPC 13156.<br />

Note: Fusicladium rhodense has a typical pseudocladosporiumlike<br />

morphology in culture, with conidial scars that are somewhat<br />

darkened <strong>and</strong> thickened.<br />

Venturia hanliniana (U. Braun & Feiler) Unter., Mycologia 89: 129.<br />

1997.<br />

Basionym: Capronia hanliniana U. Braun & Feiler, Microbiol. Res.<br />

150: 90. 1995.<br />

≡ Caproventuria hanliniana (U. Braun & Feiler) U. Braun, in Braun,<br />

A Monograph of Cercosporella, Ramularia <strong>and</strong> Allied Genera<br />

(Phytopathogenic Hyphomycetes) 2: 396. 1998.<br />

Anamorph: Fusicladium brevicatenatum (U. Braun & Feiler)<br />

Crous, U. Braun & K. Schub., comb. nov. MycoBank MB504543.<br />

Basionym: Cladophialophora brevicatenata U. Braun & Feiler,<br />

Microbiol. Res. 150: 84. 1995.<br />

≡ Pseudocladosporium brevicatenatum (U. Braun & Feiler) U. Braun,<br />

in Braun, A Monograph of Cercosporella, Ramularia <strong>and</strong> Allied Genera<br />

(Phytopathogenic Hyphomycetes) 2: 393. 1998.<br />

Venturia hystrioides (Dugan, R.G. Roberts & Hanlin) Crous & U.<br />

Braun, comb. nov. MycoBank MB504544. Fig. 28.<br />

Basionym: Capronia hystrioides Dugan, R.G. Roberts & Hanlin,<br />

Mycologia 87: 713. 1995.<br />

≡ Caproventuria hystrioides (Dugan, R.G. Roberts & Hanlin) U. Braun,<br />

in Braun, A monograph of Cercosporella, Ramularia <strong>and</strong> allied genera<br />

(Phytopathogenic Hyphomycetes). Vol. 2: 396. 1998.<br />

Anamorph: Fusicladium sp.<br />

Only the anamorph was observed on OA, PDA <strong>and</strong> SNA in culture.<br />

212


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

Fig. 27. Fusicladium rhodense (CPC 13156). A. Colony on OA. B. Conidial chains <strong>and</strong> hyphal coil. C–F. Chains of ramoconidia <strong>and</strong> conidia. Scale bar = 10 µm.<br />

Fig. 28. Venturia hystrioides (<strong>CBS</strong> 117727). A. Conidiophores giving rise to catenulate conidia. B. Ramoconidium giving rise to conidia. C–D. Conidial chains. E. Conidia <strong>and</strong><br />

conidiogenous cell with conidiogenous loci. F. Ramoconidium. G. Conidia. Scale bars = 10 µm.<br />

www.studiesinmycology.org<br />

213


Crous et al.<br />

Excluded taxa<br />

Polyscytalum fecundissimum Riess, Bot. Zeitung (Berlin) 11:<br />

138. 1853. Fig. 29.<br />

Cultural characteristics: Colonies erumpent, spreading, aerial<br />

mycelium sparse, margins smooth; colonies sienna to umber on<br />

PDA, with patches of greyish sepia; reverse chestnut-brown; on<br />

OA whitish due to moderate aerial mycelium, with diffuse umber<br />

pigment in the agar; whitish on SNA. Colonies reaching 15 mm<br />

diam on PDA after 3 wk at 25 °C in the dark.<br />

Specimen examined: Netherl<strong>and</strong>s, Schovenhorst, leaf litter of Fagus<br />

sylvatica (Fagaceae), 8 Nov. 1997, W. Gams, <strong>CBS</strong> H-6049, culture <strong>CBS</strong><br />

100506.<br />

Fig. 29. Polyscytalum fecundissimum (<strong>CBS</strong> 100506). Conidiophores giving rise to<br />

catenulate conidia. Scale bar = 10 µm.<br />

Mycelium consisting of branched, septate, smooth, guttulate,<br />

1.5–2.5 µm wide hyphae, pale brown, forming hyphal str<strong>and</strong>s.<br />

Conidiophores mostly reduced to conidiogenous cells, or if<br />

present, micronematous, consisting of a supporting cell, <strong>and</strong> single<br />

conidiogenous cell. Conidiogenous cells integrated in hyphae as<br />

lateral loci, or terminal, frequently disarticulating, subcylindrical,<br />

pale to medium brown, smooth, mono- to polyblastic, loci 1–1.5<br />

µm wide, 2.5 µm tall; conidiogenous cells subcylindrical, up to<br />

40 µm tall, <strong>and</strong> 2–2.5 µm wide. Conidia in long chains of up to<br />

60, branched or not, subcylindrical to narrowly ellipsoid, pale<br />

olivaceous to pale brown, smooth; ramoconidia 0–1(–3)-septate,<br />

15–20(–30) × 2–3(–3.5) µm; conidia 0(–1)-septate, 6–8(–12) ×<br />

2–3(–3.5) µm; hila 1–1.5 µm wide, inconspicuous to somewhat<br />

darkened, subtruncate.<br />

Cultural characteristics: Colonies erumpent, with sparse aerial<br />

mycelium on PDA, <strong>and</strong> smooth, even margins; olivaceous-grey<br />

to iron-grey (surface); reverse greenish black; on OA dark mousegrey<br />

(surface), with even, smooth margins. Colonies reaching 40<br />

mm diam after 2 wk at 25 °C in the dark; colonies fertile.<br />

Specimen examined: U.S.A., Washington, Wenatchee, on bing cherry fruit, Prunus<br />

avium cv. Bing (Rosaceae), R.G. Roberts, culture ex-type, ATCC 96019 = <strong>CBS</strong><br />

117727.<br />

Note: Dugan et al. (1995) commented that although <strong>similar</strong> to<br />

“Phaeoramularia” hachijoensis, the conidia of this species were<br />

predominantly aseptate <strong>and</strong> somewhat shorter than those described<br />

by Matsushima (1975).<br />

Notes: Polyscytalum fecundissimum is the type species of the <strong>genus</strong><br />

Polyscytalum. Several isolates of this species were investigated<br />

here to determine if Polyscytalum would be available for taxa that<br />

have a pseudocladosporium-like morphology. <strong>The</strong> clustering of<br />

<strong>CBS</strong> 681.74 within the Venturiaceae was surprising. However, this<br />

culture proved to be sterile, <strong>and</strong> therefore its identity could not be<br />

confirmed.<br />

Isolate <strong>CBS</strong> 109882 sporulated profusely. Colonies were greyolivaceous<br />

with olivaceous margins on PDA; conidiophores pale,<br />

<strong>and</strong> not dark brown as depicted for Polyscytalum in Ellis (1971);<br />

conidial chains were greenish yellow in mass, <strong>and</strong> pale olivaceousgreen<br />

under the dissecting microscope, somewhat roughened,<br />

polyblastic; on ITS sequence this isolate is identical to U57492,<br />

Cistella acuum (Alb. & Schwein.) Svrček (Helotiales), but the latter<br />

species should have a phialidic anamorph, so it is possible that<br />

this GenBank sequence is incorrect. <strong>The</strong> identity of <strong>CBS</strong> 109882<br />

therefore remains unresolved.<br />

Although isolate <strong>CBS</strong> 100506 is poorly sporulating, illustrations<br />

made in vitro when it was collected show this isolate to be<br />

authentic for the species <strong>and</strong> the <strong>genus</strong> Polyscytalum. Based on<br />

its LSU sequence, it is allied to Phlogicylindrium eucalypti Crous,<br />

Summerb. & Summerell (<strong>CBS</strong> 120080; Summerell et al. 2006), <strong>and</strong><br />

is therefore unrelated to the Venturiaceae.<br />

Zeloasperisporium R.F. Castañeda, Mycotaxon 60: 285. 1996,<br />

emend.<br />

Hyphomycetes. Mycelium mostly superficial, hyphae septate, brown<br />

to olivaceous. Hyphopodia absent. Conidiophores differentiated,<br />

mononematous, erect, aseptate or septate, brown to olivaceous.<br />

Conidiogenous cells integrated, terminal, proliferation sympodial,<br />

polyblastic, with subdenticulate, somewhat thickened <strong>and</strong> darkened<br />

scars. Conidia solitary, fusiform to obclavate or cylindrical, septate,<br />

asperulate to verrucose, olivaceous to brown, tips always hyaline,<br />

thinner-walled <strong>and</strong> smooth, forming mucoid app<strong>and</strong>ages, often only<br />

visible as a thickened frill. Synanamorph present, micronematous.<br />

Conidiogenous cells short cylindrical, antenna or hyphopodiumlike,<br />

phialidic, colarette sometimes present, aseptate, subhyaline.<br />

Conidia solitary, obovoid, ellipsoid, aseptate, brown to olivaceous,<br />

verruculose.<br />

Zeloasperisporium hyphopodioides R.F. Castañeda, Mycotaxon<br />

60: 285. 1996. Fig. 30.<br />

In vitro on OA: Mycelium internal to superficial, unbranched to<br />

sparingly branched, 1.5–3 µm wide, loosely septate, septa almost<br />

invisible, pale brown, smooth to asperulate, minutely verruculose,<br />

walls unthickened, sometimes inflated at the base of conidiophores.<br />

214


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

Fig. 30. Zeloasperisporium hyphopodioides (<strong>CBS</strong> 218.95). A–B. Conidiogenous cells. C. Conidia with apical mucoid caps. D. Conidiogenous cell with sympodial proliferation.<br />

E–G. Conidiogenous cells of micronematous synanamorph. H. Conidia, <strong>and</strong> microconidia of synanamorph. Scale bars = 10 µm.<br />

Conidiophores macronematous, arising usually laterally from<br />

plagiotropous hyphae, erect, straight, subcylindrical or conical,<br />

not geniculate, usually unbranched, rarely branched, 13–45 ×<br />

3–4(–5) µm, slightly to distinctly attenuated towards the apex,<br />

tapered, aseptate, rarely with a single septum, pale brown to pale<br />

medium brown, smooth or minutely verruculose, walls unthickened,<br />

often somewhat constricted near the base. Conidiogenous cells<br />

integrated or conidiophores usually reduced to conidiogenous<br />

cells, subcylindrical to conical, proliferation sympodial, with a single<br />

or several subdenticulate to denticulate conidiogenous loci mostly<br />

crowded at or towards the apex, protuberant, truncate, 0.8–1.2 µm<br />

wide, thickened <strong>and</strong> darkened-refractive. Conidia solitary, straight<br />

to curved, ellipsoid, fusiform to obclavate, distinctly tapered towards<br />

the apex, apiculate, (12–)15–32 × 3.5–5.5 µm, (0–)1–2(–3)-septate,<br />

mainly 1-septate, usually constricted at the septa, pale brown to<br />

pale medium brown, asperulate to verruculose, walls unthickened or<br />

almost so, tips always hyaline, thinner-walled <strong>and</strong> smooth, forming<br />

mucoid appendages, often only visible as a thickened frill, base<br />

somewhat rounded or slightly bulbous, hila often situated on short<br />

peg-like prolongations, truncate, 0.8–1(–1.2) µm wide, thickened,<br />

darkened-refractive; microcyclic conidiogenesis occurring, conidia<br />

forming secondary conidiophores.<br />

Synanamorph micronematous. Conidiophores reduced<br />

to conidiogenous cells, numerous, occurring as short lateral<br />

prolongations of hyphae, antenna or telescope-like, cylindrical,<br />

www.studiesinmycology.org<br />

unbranched, conidiogenesis unclear, at times appearing<br />

phialidic, or having one to two apical scars; up to 5 µm long,<br />

1–1.5 µm wide, aseptate, subhyaline, smooth. Conidia of the<br />

micronematous anamorph quite different from the conidia formed<br />

by the macronematous conidiophores, solitary, obovoid, ellipsoid<br />

to somewhat fusiform, 5–9 × 2.5–3 µm, aseptate, pale to pale<br />

medium brown, verruculose, somewhat attenuated towards the<br />

base, hila flat, unthickened to somewhat thickened, appearing to<br />

have the ability to form a slime appendage at the apex.<br />

Cultural characteristics: Colonies on OA iron-grey to olivaceous due<br />

to abundant sporulation (surface); reverse black, velvety; margin<br />

regular to undulate, feathery; aerial mycelium absent or sparse,<br />

sporulation profuse.<br />

Specimen examined: Cuba, isolated from air, 2 Oct. 1994, R.F. Castañeda, INIFAT<br />

C94/114, holotype, <strong>CBS</strong>-H 5624, H-5639, isotypes, culture ex-type <strong>CBS</strong> 218.95 =<br />

INIFAT C94/114 = MUCL 39155 = IMI 367520.<br />

Notes: Within the course of the recent phylogenetic studies<br />

in Herpotrichiellaceae <strong>and</strong> Venturiaceae the type culture of<br />

Zeloasperisporium hyphopodioides has been included since it was<br />

deposited at the <strong>CBS</strong> as “Fusicladium hyphopodioides”. When<br />

the culture was re-examined, the described short appressoriumlike,<br />

inflated hyphopodia with slightly warted to lobed apices<br />

(Castañeda et al. 1996) could be recognised as conidiogenous<br />

cells of a synanamorph forming a second conidial type. In addition,<br />

215


Crous et al.<br />

the conidial tips are hyaline, unthickened <strong>and</strong> smooth, <strong>and</strong> have the<br />

ability to form mucoid appendages that are often only visible as a<br />

thickened frill. <strong>The</strong>se two features, viz., the synanamorph <strong>and</strong> the<br />

conidia with mucoid appendages, easily distinguish this <strong>genus</strong> from<br />

morphologically <strong>similar</strong> genera such as Fusicladium, Asperisporium<br />

Maubl., <strong>and</strong> Passalora Fr. Phylogenetically Zeloasperisporium<br />

clusters basal to the Venturiaceae.<br />

Discussion<br />

<strong>The</strong> present paper was initiated to clarify the status of<br />

Cladophialophora <strong>and</strong> Pseudocladosporium spp., which appear<br />

morphologically <strong>similar</strong>. Confusion occurs when strains with this<br />

morphology are identified based solely on microscopic <strong>and</strong> cultural<br />

characteristics. <strong>The</strong> results clarify that Cladophialophora is allied to<br />

the Herpotrichiellaceae <strong>and</strong> Pseudocladosporium (= Fusicladium)<br />

to the Pleosporales (Dothideomycetes). <strong>The</strong> plant-pathogenic<br />

Cladophialophora species compose a separate clade within the<br />

order (Fig. 1). Another, somewhat remote chaetothyrialean clade<br />

contains extremotolerant, rock-inhabiting species around the<br />

<strong>genus</strong> Coniosporium Link (Cluster 5 of Haase et al. 1999). Both<br />

clades are significantly distinct from the prevalently hyperparasitic<br />

or oligotrophic, frequently opportunistic species of the remainder of<br />

the order (Fig. 1). This remainder includes all Capronia teleomorphs<br />

sequenced to date, <strong>and</strong> is thus likely to represent the family<br />

Herpotrichiellaceae. <strong>The</strong> ecological trends in each of the main<br />

clades of Chaetothyriales are thus quite different (Braun 1998).<br />

Several novelties are introduced within the preponderantly plantassociated<br />

clade of Chaetothyriales, including two new species<br />

associated with leaf spots. Cladophialophora is distinguished from<br />

Polyscytalum, which clusters outside the Herpotrichiellaceae, <strong>and</strong><br />

appears allied to Phlogicylindrium Crous, Summerb. & Summerell,<br />

a recently introduced <strong>genus</strong> for species occurring on Eucalyptus<br />

leaves (Summerell et al. 2006). Surprisingly Heteroconium<br />

chaetospira clusters in the Herpotrichiellaceae, <strong>and</strong> is placed<br />

in Cladophialophora as a distinctively pigmented member of the<br />

<strong>genus</strong>. Some species of Cladophialophora <strong>and</strong> Exophiala are<br />

newly described from a range of substrates such as fruit juices,<br />

drinking water <strong>and</strong> leaf litter, revealing the potential of these<br />

materials as ecological sources of inoculum for taxa associated<br />

with opportunistic human <strong>and</strong> animal infections.<br />

Furthermore, Pseudocladosporium belongs to the Venturiaceae,<br />

<strong>and</strong> is best treated as a synonym of Fusicladium, along with other<br />

genera as proposed by Schubert et al. (2003) <strong>and</strong> Beck et al. (2005).<br />

Although numerous isolates of the Venturiaceae were included for<br />

study, it was surprising to find relatively little variation within the<br />

family, suggesting that previously proposed teleomorph genera<br />

such as Apiosporina, Metacoleroa <strong>and</strong> Caproventuria should be<br />

best treated as synonyms of Venturia. <strong>The</strong> Venturiaceae is further<br />

extended with the inclusion of a novel sister clade of hyphomycetes<br />

with a pseudocladosporium-like morphology, which are also<br />

referred to as Fusicladium, thus widening the generic concept of<br />

the latter to encompass all pseudocladosporium-like anamorphs<br />

within the family. Some species assigned to Anungitopsis proved<br />

to cluster within the Venturiaceae, but the type species of the<br />

latter <strong>genus</strong>, A. speciosa, clustered elsewhere <strong>and</strong> possesses<br />

distinct conidiogenous loci, i.e., Anungitopsis cannot be reduced<br />

to synonymy with Fusicladium. <strong>The</strong> anamorphs of this sister clade<br />

of the main Venturia clade are morphologically rather close to<br />

taxa assigned to Anungitea. However, species of Anungitea <strong>and</strong><br />

Fusicladium are morphologically barely distinguishable (Schubert<br />

et al. 2003), but the true affinity of Anungitea depends on its type<br />

species of which cultures <strong>and</strong> sequence data are not yet available.<br />

Several anamorph genera with divergent morphologies were<br />

found to cluster together, suggesting that these are either different<br />

synanamorphs of the same teleomorph <strong>genus</strong>, or that they may<br />

represent cryptic clades that will diverge further once additional<br />

species are added in future studies. Although the Herpotrichiellaceae<br />

appeared to represent quite a diverse assembledge of morphotypes,<br />

the Venturiaceae were again surprisingly uniform.<br />

Acknowledgements<br />

Several colleagues from different countries provided material without which this work<br />

would not have been possible. We thank M. Vermaas for preparing the photographic<br />

plates, H.-J. Schroers for some gDNA <strong>and</strong> sequences used in this work, <strong>and</strong> A.<br />

van Iperen for maintaining the cultures. K. Schubert was financially supported by<br />

a Synthesys grant (No. 2559) which is gratefully acknowledged. R.C. Summerbell<br />

is thanked for his comments on an earlier version of the script. L. Hutchison is<br />

acknowledged for collecting Apiosporina collinsii for us.<br />

References<br />

Barr ME (1987). Prodomus to class Loculoascomycetes. University of Massachusetts,<br />

Amherst, Massachusetts.<br />

Beck A, Ritschel A, Schubert K, Braun U, Triebel D (2005). Phylogenetic relationships<br />

of the anamorphic <strong>genus</strong> Fusicladium s. lat. as inferred by ITS nrDNA data.<br />

Mycological Progress 4: 111–116.<br />

Braun U (1998). A monograph of Cercosporella, Ramularia <strong>and</strong> allied genera<br />

(Phytopathogenic Hyphomycetes). Vol. 2. IHW-Verlag, Eching.<br />

Braun U, Crous PW, Dugan F, Groenewald JZ, Hoog GS de (2003). Phylogeny <strong>and</strong><br />

taxonomy of cladosporium-like hyphomycetes, including Davidiella gen. nov.,<br />

the teleomorph of <strong>Cladosporium</strong> s.str. Mycological Progress 2: 3–18.<br />

Braun U, Feiler U (1995). Cladophialophora <strong>and</strong> its teleomorph. Microbiological<br />

Research 150: 81–91.<br />

Cai L, McKenzie EHC, Hyde KD (2004). New species of Cordana <strong>and</strong> Spadicoides<br />

from decaying bamboo culms in China. Sydowia 56: 222–228.<br />

Castañeda RF, Fabré DE, Parra MP, Perez M, Guarro J, Cano J (1996). Some<br />

airborne conidial fungi from Cuba. Mycotaxon 60: 283–290.<br />

Castañeda RF, Kendrick B (1990a). Conidial fungi from Cuba I. University of<br />

Waterloo Biological Series 32: 1–53.<br />

Castañeda RF, Kendrick B (1990b). Conidial fungi from Cuba II. University of<br />

Waterloo Biological Series 33: 1–61.<br />

Crous PW (2002). Adhering to good cultural practice (GCP). Mycological Research<br />

106: 1378–1379.<br />

Crous PW, Braun U, Schubert K, Groenewald JZ (2007a). Delimiting <strong>Cladosporium</strong><br />

from morphologically <strong>similar</strong> genera. Studies in Mycology 58: 33–56.<br />

Crous PW, Denman S, Taylor JE, Swart L, Palm ME (2004). Cultivation <strong>and</strong> diseases<br />

of Proteaceae: Leucadendron, Leucospermum <strong>and</strong> Protea. <strong>CBS</strong> Biodiversity<br />

Series 2: 1–228.<br />

Crous PW, Groenewald JZ, Wingfield MJ (2006b). Heteroconium eucalypti. Fungal<br />

Planet No. 10. (www.fungalplanet.org).<br />

Crous PW, Mohammed C, Glen M, Verkley GJM, Groenewald JZ (2007b).<br />

Eucalyptus microfungi known from culture. 3. Eucasphaeria <strong>and</strong> Sympoventuria<br />

genera nova, <strong>and</strong> new species of Furcaspora, Harknessia, Heteroconium <strong>and</strong><br />

Phacidiella. Fungal Diversity 25: 19–36.<br />

Crous PW, Slippers B, Wingfield MJ, Rheeder J, Marasas WFO, Phillips AJL, Alves<br />

A, Burgess T, Barber P, Groenewald JZ (2006a). Phylogenetic lineages in the<br />

Botryosphaeriaceae. Studies in Mycology 55: 235–253.<br />

Decock C, Delgado-Rodriguez G, Buchet S, Seng JM (2003). A new species <strong>and</strong><br />

three new combinations in Cyphellophora, with a note on the taxonomic affinities<br />

of the <strong>genus</strong>, <strong>and</strong> its relation to Kumbhamaya <strong>and</strong> Pseudomicrodochium.<br />

Antonie van Leeuwenhoek 84: 209–216.<br />

Dugan FM, Roberts RG, Hanlin RT (1995). New <strong>and</strong> rare fungi from cherry fruit.<br />

Mycologia 87: 713–718.<br />

Ellis MB (1971). Dematiaceous hyphomycetes. Commonwealth Mycological<br />

Institute, Kew.<br />

Ellis MB (1976). More dematiacous hyphomycetes. Commonwealth Mycological<br />

Institute, Kew.<br />

216


Herpotrichiellaceae <strong>and</strong> Venturiaceae<br />

Gams W, Verkley GJM, Crous PW (2007). <strong>CBS</strong> course of mycology, 5 th ed.<br />

Centraalbureau voor Schimmelcultures, Utrecht.<br />

Haase G, Sonntag L, Melzer-Krick B, Hoog GS de (1999). Phylogenetic inference<br />

by SSU gene analysis of members of the Herpotrichiellaceae, with special<br />

reference to human pathogenic species. Studies in Mycology 43: 80–97.<br />

Ho MH-M, Castañeda RF, Dugan FM, Jong SC (1999). <strong>Cladosporium</strong> <strong>and</strong><br />

Cladophialophora in culture: descriptions <strong>and</strong> an exp<strong>and</strong>ed key. Mycotaxon<br />

72: 115–157.<br />

Honbo S, Padhy AA, Ajello L (1984). <strong>The</strong> relationship of <strong>Cladosporium</strong> carrionii to<br />

Cladophialophora ajelloi. Sabouraudia 22: 209–218.<br />

Hoog GS de (1985). Taxonomy of the Dactylaria complex, IV. Dactylaria, Neta,<br />

Subulispora <strong>and</strong> Scolecobasidium. Studies in Mycology 26: 1–60.<br />

Hoog GS de, Gerrits van den Ende AHG (1998). Molecular diagnostics of clinical<br />

strains of filamentous Basidiomycetes. Mycoses 41: 183–189.<br />

Hoog GS de, Guarro, J, Gené J, Figueras MJ (2000). Atlas of clinical fungi, 2 nd<br />

ed. Centraalbureau voor Schimmelcultures, Utrecht <strong>and</strong> Universitat Rovira I<br />

Virgili, Reus.<br />

Hoog GS de, Nishikaku AS, Fern<strong>and</strong>ez Zeppenfeldt G, Padín-González C, Burger<br />

E, Badali H, Gerrits van den Ende AHG (2007). Molecular analysis <strong>and</strong><br />

pathogenicity of the Cladophialophora carrionii complex, with the description<br />

of a novel species. Studies in Mycology 58: 219–234.<br />

Hoog GS de, Takeo K, Göttlich E, Nishimura K, Miyaji M (1995). A human isolate of<br />

Exophiala (Wangiella) dermatitidis forming a catenate synanamorph that links<br />

the genera Exophiala <strong>and</strong> Cladophialophora. Journal of Medical <strong>and</strong> Veterinary<br />

Mycology 33: 355–358<br />

Jacob M, Bhat DJ (2000). Two new endophytic fungi from India. Cryptogamie<br />

Mycologie 21: 81–88.<br />

Kirk PM (1982). New or interesting microfungi. IV. Dematiaceous hyphomycetes<br />

from Devon. Transactions of the British Mycological Society 78: 55–74.<br />

Kirk PM (1983). New or interesting microfungi. IX. Dematiaceous hyphomycetes<br />

from Esher Common. Transactions of the British Mycological Society 80:<br />

449–467.<br />

Levin TP, Baty DE, Fekete T, Truant AL, Suh B (2004). Cladophialophora bantiana<br />

brain abscess in a solid-organ transplant recipient: case report <strong>and</strong> review of<br />

the literature. Journal of Clinical Microbiology 42: 4374–4378.<br />

Madan V, Bisset D, Harris P, Howard S, Beck MH (2006). Phaeohyphomycosis<br />

caused by Exophiala salmonis. British Journal of Dermatology 155: 1082–<br />

1084.<br />

Marvanová L, Laichmanová M (2007). Subilispora biappendiculata, anamorph sp.<br />

nov. from Borneo (Malaysia) <strong>and</strong> a review of the <strong>genus</strong>. Fungal Diversity 26:<br />

241–256.<br />

Matos T, Hoog GS de, Boer AG de, Crom I de, Haase G (2003). High prevalence of<br />

the neurotropic black yeast Exophiala (Wangiella) dermatitidis in steam baths.<br />

Mycoses 45: 373–377.<br />

Matsushima T (1975). Icones Microfungorum a Matsushima Lectorum. Kobe.<br />

Moncalvo J-M, Rehner SA, Vilgalys R (1993). Systematics of Lyophyllum section<br />

Difformia based on evidence from culture studies <strong>and</strong> ribosomal DNA<br />

sequences. Mycologia 85: 788–794.<br />

Partridge EC, Morgan-Jones G (2003). Notes on hyphomycetes. XC.<br />

Fusicladosporium, a new <strong>genus</strong> for cladosporium-like anamorphs of Venturia.<br />

Mycotaxon 85: 357–370.<br />

Petrak F (1949). Neue Hyphomyzeten-Gattungen aus Ekuador. Sydowia 3: 259–<br />

266.<br />

Rayner RW (1970). A mycological colour chart. CMI <strong>and</strong> British Mycological Society.<br />

Kew.<br />

Rehner SA, Samuels GJ (1994). Taxonomy <strong>and</strong> phylogeny of Gliocladium analysed<br />

from nuclear large subunit ribosomal DNA sequences. Mycological Research<br />

98: 625–634.<br />

Richards RH, Holliman A, Helgason S (1978). Exophiala salmonis infection in<br />

Atlantic salmon Salmo salar L. Journal of Fish Diseases 1: 357–368.<br />

Saccardo PA (1877). Michelia Commentarium Mycologicum Fungos in Primis<br />

Italicos Illustrans. Volume 1: 1–116. Padua, Italy.<br />

Schubert K, Braun U (2002a). Fusicladium scillae. IMI Descriptions of Fungi <strong>and</strong><br />

Bacteria 152: 1518.<br />

Schubert K, Braun U (2002b). Fusicladium convolvularum. IMI Descriptions of Fungi<br />

<strong>and</strong> Bacteria 152: 1513.<br />

Schubert K, Ritschel A, Braun U (2003). A monograph of Fusicladium s. lat.<br />

(hyphomycetes). Schlechtendalia 9: 1–132.<br />

Sivanesan A (1984). <strong>The</strong> bitunicate ascomycetes <strong>and</strong> their anamorphs. Cramer<br />

Verlag, Vaduz.<br />

Summerell BA, Groenewald JZ, Carnegie AJ, Summerbell RC, Crous PW (2006).<br />

Eucalyptus microfungi known from culture. 2. Alysidiella, Fusculina <strong>and</strong><br />

Phlogicylindrium genera nova, with notes on some other poorly known taxa.<br />

Fungal Diversity 23: 323–350.<br />

Sutton BC (1973). Hyphomycetes from Manitoba <strong>and</strong> Saskatchewan, Canada.<br />

Mycological Papers 132: 1–143.<br />

Sutton BC (1975). Hyphomycetes on cupules of Castanea sativa. Transactions of<br />

the British Mycological Society 64: 405–426.<br />

Untereiner WA (1997). Taxonomy of selected members of the ascomycete <strong>genus</strong><br />

Capronia with notes on anamorph-teleomorph connections. Mycologia 89:<br />

120–131.<br />

Untereiner WA (2000). Capronia <strong>and</strong> its anamorphs: exploring the value<br />

of morphological <strong>and</strong> molecular characters in the systematics of the<br />

Herpotrichiellaceae. Studies in Mycology 45: 141–149.<br />

Varghese KIM, Rao VG (1979). Forest microfungi – I. Subramaniomyces, a new<br />

<strong>genus</strong> of hyphomycetes. Kavaka 7: 83–85.<br />

Vilgalys R, Hester M (1990). Rapid genetic identification <strong>and</strong> mapping of<br />

enzymatically amplified ribosomal DNA from several Cryptococcus species.<br />

Journal of Bacteriology 172: 4238–4246.<br />

White TJ, Bruns T, Taylor J (1990). Amplification <strong>and</strong> direct sequencing of fungal<br />

ribosomal RNA genes for phylognetics. In: A Guide to Molecular Methods <strong>and</strong><br />

Applications (Innis MA, Gelf<strong>and</strong> DH, Sninsky JJ, White JW, eds). Academic<br />

Press, New York: 315–322.<br />

Zeng J-S, Sutton DA, Fothergill AW, Rinaldi MG, Harrak MJ, Hoog GS de (2007).<br />

Spectrum of clinically relevant Exophiala species in the U.S.A. Journal of<br />

Clinical Microbiology: in press.<br />

www.studiesinmycology.org<br />

217


available online at www.studiesinmycology.org<br />

doi:10.3114/sim.2007.58.08<br />

Studies in Mycology 58: 219–234. 2007.<br />

Molecular analysis <strong>and</strong> pathogenicity of the Cladophialophora carrionii complex,<br />

with the description of a novel species<br />

G.S. de Hoog 1,2* , A.S. Nishikaku 3 , G. Fern<strong>and</strong>ez-Zeppenfeldt 4 , C. Padín-González 4 , E. Burger 3 , H. Badali 1 , N. Richard-Yegres 4 <strong>and</strong> A.H.G.<br />

Gerrits van den Ende 1<br />

1<br />

<strong>CBS</strong> Fungal Biodiversity Centre, Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; 2 Institute for Biodiversity <strong>and</strong> Ecosystem Dynamics, University of Amsterdam, Amsterdam, <strong>The</strong> Netherl<strong>and</strong>s;<br />

3<br />

Laboratory of Immunopathology of Mycosis, Department of Immunology, São Paulo University, São Paulo, Brazil; 4 National Experimental University “Francisco de Mir<strong>and</strong>a”<br />

(UNEFM), Coro, Venezuela<br />

*Correspondence: Sybren de Hoog, s.hoog@cbs.knaw.nl<br />

Abstract: Cladophialophora carrionii is one of the four major etiologic agents of human chromoblastomycosis in semi-arid climates. This species was studied using sequence<br />

data of the internal transcribed spacer region of rDNA, the partial β-tubulin gene <strong>and</strong> an intron in the translation elongation factor 1-alpha gene, in addition to morphology.<br />

With all genes a clear bipartition was observed, which corresponded with minute differences in conidiophore morphology. A new species, C. yegresii, was introduced, which<br />

appeared to be, in contrast to C. carrionii, associated with living cactus plants. All strains from humans, <strong>and</strong> a few isolates from dead cactus debris, belonged to C. carrionii, for<br />

which a lectotype was designated. Artificial inoculation of cactus plants grown from seeds in the greenhouse showed that both fungi are able to persist in cactus tissue. When<br />

reaching the spines they produce cells that morphologically resemble the muriform cells known as the “invasive form” in chromoblastomycosis. <strong>The</strong> tested clinical strain of C.<br />

carrionii proved to be more virulent in cactus than the environmental strain of C. yegresii that originated from the same species of cactus, Stenocereus griseus. <strong>The</strong> muriform<br />

cell expressed in cactus spines can be regarded as the extremotolerant survival phase, <strong>and</strong> is likely to play an essential role in the natural life cycle of these organisms.<br />

Taxonomic novelty: Cladophialophora yegresii de Hoog, sp. nov.<br />

Key words: Cactus, chromoblastomycosis, Cladophialophora, endophyte, extremotolerance, phylogeny, taxonomy.<br />

INTRODUCTION<br />

Cladophialophora carrionii (Trejos) de Hoog, Kwon-Chung &<br />

McGinnis is one of the most frequent etiologic agents of human<br />

chromoblastomycosis, a chronic cutaneous disease characterised<br />

by verrucose skin lesions eventually leading to emerging, cauliflowerlike<br />

eruptions. <strong>The</strong> species is particularly observed in arid <strong>and</strong> semiarid<br />

climates of e.g. South <strong>and</strong> Central America (Lavelle 1980) <strong>and</strong><br />

Australia (Trejos 1954, Riddley 1957). <strong>The</strong> current hypothesis is that<br />

patients suffering from chromoblastomycosis are rural workers who<br />

acquire the infection after being pricked by cactus thorns or splinters<br />

(Rubin et al. 1991, Fernández-Zeppenfeldt et al. 1994). A classical<br />

case reported by O’Daly (1943) concerns traumatic inoculation with<br />

thorns of “guazábara” (Opuntia caribaea), a common xerophilic<br />

plant in semi-arid Venezuela. This hypothesised traumatic route of<br />

infection was later supported by Richard-Yegres & Yegres (Richard-<br />

Yegres & Yegres 1987; strain SR3 = <strong>CBS</strong> 863.96) <strong>and</strong> Fernández-<br />

Zeppenfeldt et al. (1994), who isolated strains from Prosopis<br />

juliflora litter. Cladophialophora Borelli has also been detected in<br />

association with spines of the common xerophyte Aloe vera <strong>and</strong> of<br />

the Cactaceae: Opuntia caribaeae, O. caracasana, Stenocereus<br />

griseus <strong>and</strong> Cereus lanuginosus (Borelli 1972, Yegres et al. 1996).<br />

Thorny American cacti are important components of the xerophyte<br />

flora of the arid climate of our study area in Falcon State, Venezuela<br />

(Richard-Yegres et al. 1992). Muriform cells are produced in<br />

human skin <strong>and</strong> represent the supposed pathogenic invasive form<br />

of fungi causing chromoblastomycosis (Mendoza et al. 1993).<br />

Early experiments involving the inoculation of several species of<br />

cold-blooded animals have shown the abundant production of the<br />

characteristic muriform cells in vivo (Trejos 1953).<br />

A <strong>similar</strong> plant origin of chromoblastomycosis has been<br />

supposed for a related agent of chromoblastomycosis, Fonsecaea<br />

pedrosoi (Brumpt) Negroni. Marques et al. (2006) isolated this<br />

species from the shells of Babassu coconuts (Orbignya phalerata).<br />

<strong>The</strong> habit of local people to sit on these shells might explain the<br />

frequent occurrence of lesions on the buttocks (Silva et al. 1995).<br />

Salgado et al. (2004) found the species on the thorns of a Mimosa<br />

pudica plant which a patient could identify as the source of traumatic<br />

onset of his chromoblastomycosis.<br />

Recently, with the development of molecular tools for species<br />

identification, doubt has arisen about the correctness of this<br />

supposed route of infection. <strong>The</strong> question whether environmental<br />

<strong>and</strong> clinical strains represent exactly the same species needs to be<br />

re-determined. In order to establish this for C. carrionii-associated<br />

chromoblastomycosis, reference strains from the <strong>CBS</strong> culture<br />

collection, supplemented with a large set of strains from semiarid<br />

Venezuela, have been verified using molecular tools that are<br />

currently routinely employed to answer taxonomic questions in<br />

black yeasts <strong>and</strong> their filamentous relatives (de Hoog et al. 2003),<br />

particularly the internal transcribed spacer (ITS) region of rDNA,<br />

the partial β-tubulin gene (BT2), <strong>and</strong> an intron in the translation<br />

elongation factor 1-alpha (EF1). In addition, a series of three<br />

experiments has been conducted concerning inoculation into<br />

<strong>and</strong> superficial application onto germlings of Stenocereus griseus<br />

obtained by cultivation in vitro, mature plants of S. griseus from the<br />

wild, <strong>and</strong> in spines of S. griseus collected in the semi-arid area of<br />

study. Our aim is to reveal the role of the cactus S. griseus in the<br />

life cycle of its associated Cladophialophora spp., <strong>and</strong> to determine<br />

whether a link could be made to C. carrionii for obtaining a better<br />

underst<strong>and</strong>ing of human chromoblastomycosis.<br />

MATERIALS AND METHODS<br />

Fungal strains <strong>and</strong> morphology<br />

Strains studied are listed in Table 1. This list comprises strains which<br />

have morphologically been identified as C. carrionii. Reference<br />

strains from the <strong>CBS</strong> culture collection, as well as fresh isolates<br />

from patients <strong>and</strong> the environment have been included. Strains<br />

were lyophilised <strong>and</strong> stored in liquid nitrogen soon after deposit at<br />

<strong>CBS</strong>. Stock cultures for transient working collections were grown<br />

on slants of 2 % malt extract agar (MEA) <strong>and</strong> oatmeal agar (OA) at<br />

24 °C. For morphological observation, slide cultures were made of<br />

strains grown on potato-dextrose agar (PDA) (de Hoog et al. 2000)<br />

<strong>and</strong> mounted in lactophenol cotton blue.<br />

219


De Hoog et al.<br />

Table 1. Isolation data of Cladophialophora strains examined.<br />

Name <strong>CBS</strong> nr. Other reference(s) 1 GenBank mtDNA*<br />

(Kawasaki<br />

ITS, BT2, EF1 et al. 1993)<br />

Source [human: duration, localization, sex, age<br />

(Pérez-Blanco et al. 2003)].<br />

Geography<br />

A / I / 1: C. carrionii<br />

117904 UNEFM 0004-02 = dH 14480 EU137281, –, – Chromoblastomycosis; 14 y; hip, thigh, leg; male 38 Falcon State, Venezuela<br />

117891 UNEFM 0002-00 = dH 14475 EU137278, –, EU137222 Chromoblastomycosis; 1 y; male 62 Falcon State, Venezuela<br />

117906 UNEFM 0014-96 = dH 14504 EU137288, EU137171, EU137231 Chromoblastomycosis; 0.5 y; h<strong>and</strong>; male 45 Falcon State, Venezuela<br />

117897 UNEFM 0011-03 = dH 14497 EU137314, –, EU137254 Chromoblastomycosis; 0.5 y; h<strong>and</strong>; male 42 Falcon State, Venezuela<br />

859,96 UNEFM 9617 = dH 10703 EU137295, EU137178, EU137237 Dry plant debris, arid zone Falcon State, Venezuela<br />

117898 UNEFM 0010-98 = dH 14496 EU137308, –, EU137246 Chromoblastomycosis; 20 y; h<strong>and</strong>; female 59 Falcon State, Venezuela<br />

117889 UNEFM 0003-04 = dH 14478 –, EU137190, – Chromoblastomycosis; 20 y; thigh, leg; female 78 Falcon State, Venezuela<br />

114392 UNEFM 82267 = dH 13261 EU137267, EU137150, EU137211 Chromoblastomycosis; leg; female Falcon State, Venezuela<br />

114394 UNEFM 9803 = dH 13263 EU137307, –, EU137245 Chromoblastomycosis; h<strong>and</strong>; male 22 Falcon State, Venezuela<br />

114396 UNEFM 2001/1 = dH 13265 EU137269, EU137152, EU137213 Chromoblastomycosis; arm; male 35 Falcon State, Venezuela<br />

114399 UNEFM 2003/2 = dH 13268 EU137272, EU137155, EU137216 Chromoblastomycosis; arm; female 64 Falcon State, Venezuela<br />

114401 UNEFM 9901 = dH 13270 EU137274, EU137157, EU137218 Chromoblastomycosis; arm; female 40 Falcon State, Venezuela<br />

114402** UNEFM 9902 = dH 13271 EU137275, EU137158, EU137219 Chromoblastomycosis; arm; female 40 Falcon State, Venezuela<br />

114403 UNEFM 95195 = dH 13272 EU137276, EU137159, EU137220 Chromoblastomycosis; arm; male Falcon State, Venezuela<br />

117899 UNEFM 0010-04 = dH 14495 EU137301, EU137183, EU137241 Chromoblastomycosis; 2 y; h<strong>and</strong>; male 57 Falcon State, Venezuela<br />

117901 UNEFM 0009-03 = dH 14492 EU137312, EU137197, EU137252 Chromoblastomycosis; 8 y; arm; female 41 Falcon State, Venezuela<br />

114393 UNEFM 9801 = dH 13262 EU137268, EU137151, EU137212 Chromoblastomycosis; h<strong>and</strong>; male 72 Falcon State, Venezuela<br />

108.97** UNEFM 9501 = dH 10704 EU137306, EU137188, EU137265 Chromoblastomycosis; skin Falcon State, Venezuela<br />

114397 UNEFM 84020 = dH 13266 EU137270, EU137153, EU137214 Chromoblastomycosis; h<strong>and</strong>, arm; male 54 Falcon State, Venezuela<br />

114404 UNEFM 95656 = dH 13273 EU137311, EU137196, EU137251 Chromoblastomycosis; arm; male Falcon State, Venezuela<br />

117902 UNEFM 0008-03 = dH 14489 EU137283, EU137166, EU137226 Chromoblastomycosis; 3 y; arm; male 42 Falcon State, Venezuela<br />

117893 UNEFM 0001-00 = dH 14470 EU137316, EU137200, – Chromoblastomycosis; 2 y; knee; male 19 Falcon State, Venezuela<br />

117892 UNEFM 0001-02 = dH 14471 EU137277, EU137160, EU137221 Chromoblastomycosis; 8 y; knee; male 52 Falcon State, Venezuela<br />

117908 UNEFM 0013-04 = dH 14502 –, EU137191, – Chromoblastomycosis; 6 y; back; male 13 Falcon State, Venezuela<br />

109.97** UNEFM 9503 = dH 10706 –, –, – Chromoblastomycosis; skin Falcon State, Venezuela<br />

857.96 UNEFM 9408 = dH 10707 EU137294, EU137177, EU137236 Chromoblastomycosis; skin Falcon State, Venezuela<br />

114398 UNEFM 2003/1 = dH 13267 EU137271, EU137154, EU137215 Chromoblastomycosis; arm; female 67 Falcon State, Venezuela<br />

114400 UNEFM 2003/3 = dH 13269 EU137273, EU137156, EU137217 Chromoblastomycosis; arm; male 50 Falcon State, Venezuela<br />

117909 UNEFM 0013-00 = dH 14501 EU137287, EU137170, EU137230 Chromoblastomycosis; arm; male Falcon State, Venezuela<br />

114395 UNEFM 9802 = dH 13264 EU137299, EU137182, EU137240 Chromoblastomycosis; leg; female 22 Falcon State, Venezuela<br />

166.54 MUCL 10088 EU137290, EU137173, – Skin lesion in human Falcon State, Venezuela<br />

862.96 UNEFM 9603 = dH 10700 EU137315, EU137199, EU137255 Dry plant debris, semi-arid zone Falcon State, Venezuela<br />

863.96** IFM 41444 = UNEFM SR3 = dH 10699 AB109169 / EU137296, EU137179, EU137238 Dry spine (Opuntia caribaea) on soil, semi-arid zone Falcon State, Venezuela<br />

861.96 UNEFM 9607 = dH 10701 EU137309, EU137194, EU137249 Dry plant debris, semi-arid zone Falcon State, Venezuela<br />

117896 dH 14498 EU137285, –, EU137228 H<strong>and</strong> lesion Falcon State, Venezuela<br />

114397 UNEFM 84020 = dH 13266 EU137270, EU137153, EU137214 Chromoblastomycosis, h<strong>and</strong> <strong>and</strong> arm Falcon State, Venezuela<br />

220


Cladophialophora carrionii complex<br />

117905 dH 14505 EU137300, –, – Chromoblastomycosis, h<strong>and</strong>, male Falcon State, Venezuela<br />

117900 dH 14493 EU137284, –, EU137227 Chromoblastomycosis, h<strong>and</strong>, male Falcon State, Venezuela<br />

114392 UNEFM 82267 = dH 13261 EU137267, EU137150, EU137211 Chromoblastomycosis, leg, female Falcon State, Venezuela<br />

– FMC 248 AF397181, –, – Chromoblastomycosis Venezuela<br />

– IFM 41807 AB109175, –, – Group mt-I – Venezuela<br />

– IFM 4812 AB109168, –, – Group mt-I – Venezuela<br />

– IMTSP 690 AF397180, –, – Chromoblastomycosis Brazil<br />

410.96 UAMH 4392 = NCMH 1010 = DUKE 2403 EU137310, EU137195, EU137250 Chromoblastomycosis –<br />

163.54 EU137304, EU137186, EU137243 Chromoblastomycosis Australia<br />

117903 dH 14482 EU137282, –, EU137225 Chromoblastomycosis, forearm, male –<br />

362.70 M.J. Campos 4555 = dH 15806 EU137302, EU137184, EU137242 Human Mozambique<br />

260.83 CDC B-1352 = FMC 282 = ATCC 44535 (ex-T of C. ajelloi) EU137292, EU137175, EU137234 Group mt-I Skin lesion in human Ug<strong>and</strong>a<br />

986.96 UAMH 5717 EU137297, EU137180, – Clinical material –<br />

– IFM 4805 AB087204, –, – – –<br />

– IFM 4811 AB109178, –, – – –<br />

– IFM 41814 AB109176, –, – – –<br />

B / II / 2: C. carrionii<br />

160.54 ATCC 16264 = CDC A-835 = MUCL 40053 = IFM 4808 (ex-LT of C. carrionii) AB109177 / EU137266, EU137201, EU137210 Group mt-II Chromoblastomycosis, human Australia<br />

– Todd Pryce 200867 = dH 13218 Human Australia<br />

406.96 MRL 1114 = UAMH 4366 = dH 15847 EU137317, EU137202, EU137256 Human Queensl<strong>and</strong>, Australia<br />

100434 ATCC 32279 = dH 10745 = IP 518 = RV 16499 EU137289, EU137172, EU137232 Human Madagascar<br />

– IFM 4810 AB109170, –, – – –<br />

– IFM 41446 = DCU 606 AB109171, –, – – –<br />

C: C. carrionii<br />

– IFM 41651 AB109174, –, – Group mt-I – China<br />

– IFM 41650 AB109173, –, – Group mt-I – China<br />

– IFM 41641 AB109172, –, – Group mt-I – China<br />

– IFM 4985 AB109179, –, – – –<br />

– IFM 4986 AB109180, –, – – –<br />

D / III / 3: C. yegresii<br />

114406 UNEFM SgSR1 = dH 13275 EU137323, EU137208, EU137263 Stenocereus griseus asymptomatic plant Falcon State, Venezuela<br />

114407 UNEFM SgSR2 = dH 13276 EU137324, –, EU137264 Stenocereus griseus asymptomatic plant Falcon State, Venezuela<br />

114405** UNEFM SgSR3 = dH 13274 (ex-T of C. yegresii) EU137322, EU137209, EU137262 Stenocereus griseus asymptomatic plant Falcon State, Venezuela<br />

1 Abbreviations: ATCC = American Type Culture Collection, Manassas, U.S.A.; <strong>CBS</strong> = Centraalbureau voor Schimmelcultures, Utrecht, <strong>The</strong> Netherl<strong>and</strong>s; CDC = Centers for Disease Control <strong>and</strong> Prevention, Atlanta, U.S.A.; DCU = Department of<br />

Dermatology, School of Medicine, Chiba, Japan; dH = G.S. de Hoog working collection; FMC = Faculdade de Medicina, Caracas, Venezuela; ITMSP = Instituto de Medicina Tropical de São Paulo, São Paulo, Brazil; IFM = Research Center for<br />

Pathogenic Fungi <strong>and</strong> Microbial Toxicoses, Chiba University, Chiba, Japan; IP = Institut Pasteur, Paris, France; MUCL = Mycotheque de l’Université de Louvain, Louvain-la-Neuve, Belgium; RV = Prince Leopold Institute of Tropical Medicine, Antwerp,<br />

Belgium; UAMH = <strong>The</strong> University of Alberta Microfungus Collection <strong>and</strong> Herbarium, Edmonton, Canada; UNEFM = Universidade Nacional Experimental Francisco de Mir<strong>and</strong>a, Coro, Falcon, Venezuela.<br />

Ex-T = Type strain; ex-LT = Lectotype strain.<br />

*Type I <strong>and</strong> II groups based on mitochondrial DNA restriction fragment length polymorphism: H-1 <strong>and</strong> H-2 = restriction patterns 1 <strong>and</strong> 2, respectively, with HaeIII enzyme; M-1 <strong>and</strong> M-2 = restriction patterns 1 <strong>and</strong> 2, respectively, with MspI enzyme; S-1<br />

<strong>and</strong> S-2 = restriction patterns 1 <strong>and</strong> 2, respectively, with Sau3AI enzyme.<br />

** Used in plant <strong>and</strong> mouse inoculation experiments.<br />

www.studiesinmycology.org<br />

221


De Hoog et al.<br />

Table 2. Results from MrAic using corrected Akaike Information Criterion (AICc).<br />

Fragment/Gene Model df* lnL* AICc* wAICc*<br />

rRNA ITS TrNG 89 -21.840.556 4575.9992” 0.3080<br />

EF-1α HKYG 88 -19.392.355 41.147.171 0.5242<br />

ß-Tubulin SYMIG 90 -28.547.745 59.261.933 0.1913<br />

*df = degrees of freedom; lnL = log likelihood; AICc = corrected AIC; wAICc = weighted corrected AIC.<br />

DNA extraction<br />

Approximately 1 cm 2 mycelium of 30-d-old cultures was transferred<br />

to a 2 mL Eppendorf tube containing 300 µL TES-buffer (Tris 1.2 %<br />

w/v, Na-EDTA 0.38% w/v, SDS 2 % w/v, pH 8.0) <strong>and</strong> about 80 mg<br />

of a silica mixture (Silica gel H, Merck 7736, Darmstadt, Germany<br />

/ Kieselguhr Celite 545, Machery, Düren, Germany, 2 : 1, w/w).<br />

Cells were disrupted mechanically in a tight-fitting sterile pestle for<br />

approximately 1 min. Subsequently 200 µL TES-buffer was added,<br />

the mixture was vortexed, 10 µL proteinase K was added <strong>and</strong><br />

incubated for 10 min at 65 °C. After addition of 140 µL of 5 M NaCl<br />

<strong>and</strong> 1/10 vol CTAB 10 % (cetyltrimethylammoniumbromide) buffer,<br />

the material was incubated for 30 min at 65 °C. Subsequently<br />

700 µL SEVAG (24 : 1, chloroform : isoamylalcohol) was mixed to<br />

solution, incubated during 30 min on ice water <strong>and</strong> centrifuged for<br />

10 min at 14 000 rpm. <strong>The</strong> supernatant was transferred to a new<br />

tube with 225 µL 5 M NH 4<br />

-acetate, incubated on ice water <strong>and</strong><br />

centrifuged again for 10 min at 14 000 rpm. <strong>The</strong> supernatant was<br />

transferred to another Eppendorf tube with 0.55 vol isopropanol <strong>and</strong><br />

spun for 5 min at 14 000 rpm. Subsequently, the pellet was washed<br />

with ice cold 70 % ethanol. After drying at room temperature it was<br />

re-suspended in 48.5 µL TE buffer (Tris 0.12 % w/v, Na-EDTA 0.04<br />

% w/v) plus 1.5 µL RNAse 20 U/mL <strong>and</strong> incubated for 15–30 min<br />

at 37 °C.<br />

Sequencing <strong>and</strong> phylogenetic reconstruction<br />

Three loci, namely the internal transcribed spacers (ITS), ß-<br />

tubulin (BT2) <strong>and</strong> translation elongation factor 1-α (EF1), were<br />

sequenced. For ITS sequencing, amplification was performed<br />

with V9G (5’-TTACGTCCCTGCCCTTTGTA-3’) <strong>and</strong> LS266 (5’-<br />

GCATTCCCAAACAACTCGACTC-3’). Sequencing reactions<br />

were conducted with ITS1 <strong>and</strong> ITS4 primers (White et al.<br />

1990). For BT2 amplification <strong>and</strong> sequencing, primers Bt2a<br />

(5’-GGTAACCAAATCGGTGCTGCTTTC-3’) <strong>and</strong> Bt2b (5’-<br />

ACCCTCAGTGTAGTGACCCTTGGC-3’) were used (Glass &<br />

Donaldson 1995) <strong>and</strong> for EF1 amplification <strong>and</strong> sequencing, primers<br />

EF1-728F (5’CATCGAGAAGTTCGAGAAGG-3’) <strong>and</strong> EF1-986R<br />

(5’-TACTTGAAGGAACCCTTACC-3’) (Carbone & Kohn, 1999).<br />

Sequences were aligned in BioNumerics v. 4.5 (Applied Maths,<br />

Kortrijk, Belgium), exported <strong>and</strong> converted into Phylip interleaved<br />

format (Felsenstein 1993).<br />

Calculation of ILD (incongruence length difference) was<br />

performed in PAUP v. 4.0b10 (Swofford 2003). A combined data set<br />

of ITS, EF1 <strong>and</strong> BT2 sequences was created. Optimality criterion<br />

was set to parsimony. <strong>The</strong> total number of characters was 1 263<br />

with equal weight, while 677 characters were constant, <strong>and</strong> 396<br />

parsimony-informative. Gaps were treated as missing, <strong>and</strong> treebisection-reconnection<br />

(TBR) was used as branch-swapping<br />

algorithm. Maximum number of trees was set to 100 <strong>and</strong> left<br />

unchanged.<br />

Substitution model testing<br />

<strong>The</strong> program MrAic (www.abc.se/~nyl<strong>and</strong>er/; Nyl<strong>and</strong>er 2004)<br />

was used to select a substitution model. MrAic is a Perl script for<br />

calculating the Akaike Information Criterion (AIC), corrected Akaike<br />

Information Criterion (AICc), Bayesian Information Criterion (BIC),<br />

<strong>and</strong> Akaike weights for nucleotide substitution models <strong>and</strong> model<br />

uncertainty. Using an ML algorithm, likelihood scores under different<br />

models were estimated using Phyml (http://atgc.lirmm.fr/phyml/). All<br />

56 models implemented in Modeltest (Posada & Cr<strong>and</strong>all 1998)<br />

were evaluated. <strong>The</strong>se models were also combined with proportion<br />

of invariable sites (I) <strong>and</strong>/or gamma distribution shape parameter<br />

(G). A difference between Modeltest <strong>and</strong> MrAic is that the latter<br />

does not evaluate all models on the same, approximate topology<br />

as in PAUP (Swofford 1981). Instead, Phyml was used to try to find<br />

the maximum of the likelihood function under all models. This is<br />

necessary for finding AIC, AICc, or BIC for the models. <strong>The</strong> AICc<br />

calculation (Table 2) was used to select the right model for the ratio<br />

of parameters to characters (Nchar/Nparameters < 40; Burnham &<br />

Anderson 2002) for all loci. <strong>The</strong> substitution matrix of the models<br />

is printed next to the trees. Another advantage of using MrAic in<br />

combination with Phyml was the obtained accuracy of tree topology<br />

<strong>and</strong> the greater calculation speed (Guindon & Gascuel 2003).<br />

Population genetic analyses<br />

In order to confirm the intraspecific diversity shown in the MP trees,<br />

the number of populations in the C. carrionii complex was inferred<br />

with Structure v. 2.2 (Pritchard et al. 2000) using genotype data<br />

of the ITS regions of rRNA gene <strong>and</strong> of the partial EF1 <strong>and</strong> BT2<br />

genes. Genotypes of these three loci of 43 isolates were sorted<br />

on the basis of sequence <strong>similar</strong>ity. Structure is a model-based<br />

clustering method for using multilocus genotype data to infer<br />

population structure <strong>and</strong> assign individuals to populations. <strong>The</strong><br />

parameters were as follows: the length of burn-in period was<br />

set to 10 6 , number of MCMC repeats after burn-in 30 000; the<br />

ancestry model: admixture (individuals have mixed ancestry <strong>and</strong><br />

is recommended as starting point for most analyses). Uniform prior<br />

for ALPHA was set to 1.0 (default) <strong>and</strong> all allele frequencies were<br />

taken as independent among populations with λ set to 1.0 (default).<br />

Probability of the data (for estimating K) was also computed<br />

(Falush et al. 2003). <strong>The</strong> burn-in period length <strong>and</strong> number of<br />

MCMC repetitions after burn-in were set as 10 000 <strong>and</strong> 100 000,<br />

<strong>and</strong> admixture model <strong>and</strong> allele frequencies correlated model were<br />

chosen for analysis. <strong>The</strong> number of populations (K) was assumed<br />

from two to four.<br />

Association of multilocus genotypes was screened with the<br />

multilocus option in BioNumerics. To test for reproductive mode in<br />

each population, index of association (I A<br />

, a measure of multilocus<br />

linkage disequilibrium) was calculated with Multilocus v. 1.2.2 (www.<br />

bio.ic.ac.uk/evolve/software/multilocus). <strong>The</strong> null hypothesis for this<br />

analysis was complete panmixia. <strong>The</strong> values of I A<br />

were compared<br />

between observed <strong>and</strong> r<strong>and</strong>omised data sets. <strong>The</strong> hypothesis would<br />

be rejected when p < 0.05. Population differentiation (index: theta,<br />

θ) was also detected using the same software <strong>and</strong> a null hypothesis<br />

for this analysis is no population differentiation. When observed θ is<br />

statistically significantly different from those of r<strong>and</strong>om datasets (p<br />

< 0.05), population differentiation should be considered.<br />

222


Cladophialophora carrionii complex<br />

A reticulogram was reconstructed using T-rex (Makarenkov<br />

2001, Makarenkov & Legendre 2004) (www.labunix.uqam.ca/<br />

~makarenv/trex.html) on C. carrionii / Cladophialophora sp. <strong>The</strong><br />

program first computed a classical additive tree using one of the<br />

five available tree reconstruction algorithms. Subsequently, at each<br />

step of the procedure, a reticulation (a new edge) was chosen that<br />

minimised the least-squares or the weighted least-squares loss<br />

function; it was added to the growing reticulogram. Two statistical<br />

criteria (Q1 <strong>and</strong> Q2) were proposed to measure the gain in fit when<br />

reticulations were added. <strong>The</strong> minimum of each of these criteria<br />

may suggest a stopping rule for addition of reticulations. With<br />

HGT (horizontal gene transfer) reticulogram reconstruction option<br />

(Makarenkov 2001) the program mapped the gene tree into the<br />

species tree using the least-squares method. Horizontal transfers<br />

of the considered gene were then shown in the species tree. <strong>The</strong><br />

reticulate network was created in the ITS tree, which served as<br />

a species tree <strong>and</strong> compared with a gene tree, EF1. Degrees of<br />

recombination or horizontal gene transfer were also visualised<br />

using SplitsTree v. 4.8 software (Huson & Bryant 2006). Split<br />

decomposition (B<strong>and</strong>elt & Dress 1992) was applied on three loci of<br />

the entire C. carrionii complex. Calculation was done with default<br />

settings of characters transformation using uncorrected P-values,<br />

equal angles <strong>and</strong> optimise box iterations set to 1. Star- or brushlike<br />

trees indicate clonal development, while reticulation indicates<br />

genetic exchange.<br />

Isolation of fungi for inoculation experiments<br />

Nine plants of Stenocereus griseus, located within 50 m radius<br />

of the house of a patient with chromoblastomycosis due to strain<br />

UNEFM 9902 = <strong>CBS</strong> 114402 (C. carrionii) in Sabaneta (Mir<strong>and</strong>a,<br />

Falcón State, Venezuela), were analysed. Four fragments of<br />

approx. 2 × 3 × 1 cm were excised from each plant at brownish<br />

superficial lesions in upper branches. Sampled fragments were<br />

soaked in mineral oil for 15 min at 23 °C under agitation at 150<br />

rpm (Fernández-Zeppenfeldt et al. 1994). Subsequently four<br />

cultivations were made per sample on agar slants. Strains with<br />

cultural characteristics <strong>and</strong> morphology <strong>similar</strong> to C. carrionii (de<br />

Hoog et al. 2000) were selected. Final identification was made by<br />

sequencing of the ITS region, by determining the ability of strains<br />

to grow at 35, 37, 38 <strong>and</strong> 40 °C, <strong>and</strong> whether they could break<br />

down 20 % gelatin (Richard-Yegres & Yegres 1987, Fernández-<br />

Zeppenfeldt et al. 1994). Environmental strain UNEFM-SgSR3 =<br />

<strong>CBS</strong> 114405 (Cladophialophora sp.) <strong>and</strong> clinical strain UNEFM<br />

9902 = <strong>CBS</strong> 114402 (C. carrionii) were selected for the inoculation<br />

experiments.<br />

Inoculum preparation<br />

Approximately 1 cm 2 of a culture on Sabouraud’s glucose agar<br />

(SGA) was transferred to 50 mL YPG medium (yeast extract 0.5<br />

%, peptone 0.5 %, glucose 2 %) (de Hoog et al. 2000), shaken at<br />

150 rpm <strong>and</strong> incubated for 3 d at 23 °C (Yegres et al. 1991). Five<br />

mL aliquots of the starter culture were transferred serially every 4<br />

d to 500 mL flasks containing 100 mL synthetic medium (d-glucose<br />

2 %, KH 2<br />

PO 4<br />

0.2 %, NH 4<br />

SO 4<br />

0.1 %, urea 0.03 %, MgSO 4<br />

0.03<br />

%, CaCl 2<br />

0.003 %; pH 6.2) shaken at 150 rpm at 23 °C. After 4 d<br />

the suspensions, which were predominantly conidia, were filtered<br />

through sterile gauze, ground in 50 mL 0.85 % saline, centrifuged<br />

at 2 000 rpm, <strong>and</strong> repeatedly washed with saline until a clear<br />

supernatant was obtained. <strong>The</strong> suspensions were adjusted to 5 ×<br />

10 6 cells/mL (Yegres et al. 1991, Cermeño & Torres 1998). Inocula<br />

of 2 mL were checked for viability in lactritmel medium (de Hoog et<br />

al. 2000).<br />

www.studiesinmycology.org<br />

Experimental cactus germlings<br />

Young cactus plants (Stenocereus griseus) were obtained in the<br />

laboratory (Clausnitzer 1978) by cultivation from seeds of a single<br />

cardon fruit collected near the house of the patient infected with <strong>CBS</strong><br />

114402 in the endemic area for chromoblastomycosis in Falcon<br />

State, Venezuela. <strong>The</strong> seeds were rinsed with sterile distilled water,<br />

the contents washed by agitation for 10 min at 120 rpm in 250 mL<br />

sodium hypochlorite 4 % (v/v), <strong>and</strong> subsequently with sterile distilled<br />

water at 120 rpm for 5 min. <strong>The</strong> supernatant was decanted, 250 mL<br />

HCl 20 % was added, the seeds were incubated for 3 h, decanted<br />

<strong>and</strong> washed repeatedly with sterile distilled water. Seeds were<br />

then dried for 24 h on filter paper at 37 °C. Onset of germination<br />

was obtained by incubation of the seeds in a moist chamber on<br />

filter paper for 15 d under alternately 8 h of continuous white light<br />

(26 W) <strong>and</strong> 16 h of darkness; bud emergence was observed daily.<br />

Germlings of 1 cm in length, with green colour <strong>and</strong> having two<br />

leaves were transplanted to 128-container germinators until roots<br />

developed. <strong>The</strong> sterile substrate contained 5 parts Sogemix® <strong>and</strong><br />

1 part river soil from the region where the fruit was collected. <strong>The</strong><br />

daily light regime was as above; plants were watered every 10 d<br />

with 5 mL sterile tap water for 1 yr.<br />

Inoculation of S. griseus germlings<br />

Fungal suspensions (0.1 mL) were either injected using a syringe<br />

(13 × 0.4 mm) at a depth of approximately 5 mm into cortical<br />

tissue (Fig. 1A), or superficially applied onto (Fig. 1B) 96 r<strong>and</strong>omly<br />

selected 1-yr-old plants: 50 % using clinical strain <strong>CBS</strong> 114402<br />

(C. carrionii) <strong>and</strong> 50 % using environmental strain <strong>CBS</strong> 114405<br />

(Cladophialophora sp.). <strong>The</strong> controls were 64 plants which were<br />

treated <strong>similar</strong>ly, but using sterile saline (0.85 %). <strong>The</strong> growth<br />

chambers with inoculated plants stayed in the laboratory under the<br />

conditions specified above. From day 15 post inoculation onwards<br />

every 15 th day, six plants of each treatment were sectioned<br />

longitudinally from the apex <strong>and</strong> transversely by means of a h<strong>and</strong>held<br />

microtome, examined directly in glycerin water (25 %), <strong>and</strong><br />

cultured in lactritmel medium (Fernández-Zeppenfeldt et al. 1994).<br />

Experimental cactus plants<br />

A total of 150 whole S. griseus plants ≤ 15 cm tall <strong>and</strong> without<br />

macroscopically visible lesions, were dug from an area within a<br />

50 m radius of the house of the patient with chromoblastomycosis<br />

as specified above. Plants were transported to the laboratory <strong>and</strong><br />

transplanted individually into polyethylene bags with a capacity of<br />

1 kg, using as substrate river soil from the same area. Plants were<br />

maintained outside, directly adjacent to the laboratory to adjust<br />

at average temperatures of 32 °C <strong>and</strong> with natural daylight. <strong>The</strong>y<br />

were watered with tap water every 15 th d for a period of 6 mo.<br />

Scar formation in mature S. griseus plants<br />

For inoculation purposes, 150 sharp, wooden toothpicks 4 × 0.3 ×<br />

0.2 cm were washed <strong>and</strong> boiled for 3 min in tap water to eliminate<br />

resins (Yegres & Richard-Yegres 2002). This procedure was<br />

repeated three times. Three batches of 50 toothpicks each were<br />

kept separate in Petri dishes. Plates were incubated for 15 d at<br />

23 °C after inoculating each batch with 1 mL fungal suspension (5<br />

× 10 6 cells/mL) of either strain <strong>CBS</strong> 114405 or <strong>CBS</strong> 114402, with<br />

sterile water as control. A total of 50 r<strong>and</strong>omly selected plants were<br />

inoculated (Fig. 1C) halfway up the shaft with a toothpick colonised<br />

with <strong>CBS</strong> 114402 (C. carrionii), <strong>CBS</strong> 114405 (Cladophialophora<br />

sp.) or the control (Yegres & Richard-Yegres 2002).<br />

Starting from 2 wk post inoculation, 10 plants were r<strong>and</strong>omly<br />

chosen every 15 d, <strong>and</strong> tissue samples taken at the point of<br />

223


De Hoog et al.<br />

+ + + - -<br />

-<br />

+<br />

+<br />

A B C D<br />

Clinical C. carrionii<br />

> Environmental C. yegresii<br />

Fig. 1. Diagram of inoculation experiments with results. A. Inoculation of young cactus; B. Superficial application of young cactus; C. Traumatic application of mature cactus,<br />

with brown resulting scar; D. Superficial application of mature spines. Indications +/- refer to positive resp. negative results of re-isolated strains. Lower line: circles represent<br />

production of muriform cells, filaments represent hyphal growth.<br />

Fig. 2. Split decomposition of the C. carrionii complex using SplitsTree with uncorrected (P-value) distances. Nodes are shown only with different genotypes; hence EF1 shows<br />

the largest number of nodes. Extensive reticulation is noted in all loci.<br />

224


Cladophialophora carrionii complex<br />

0.1<br />

99<br />

<strong>CBS</strong>102227<br />

<strong>CBS</strong>83496<br />

92<br />

94<br />

85<br />

<strong>CBS</strong>25983<br />

<strong>CBS</strong>45482<br />

100<br />

<strong>CBS</strong>117901<br />

<strong>CBS</strong>117891<br />

<strong>CBS</strong>117906<br />

<strong>CBS</strong>117892<br />

<strong>CBS</strong>114403<br />

<strong>CBS</strong>117899<br />

<strong>CBS</strong>114392<br />

<strong>CBS</strong>114393<br />

<strong>CBS</strong>41096<br />

<strong>CBS</strong>117903<br />

<strong>CBS</strong>114395<br />

<strong>CBS</strong>114398<br />

<strong>CBS</strong>114404<br />

<strong>CBS</strong>117900<br />

<strong>CBS</strong>117909<br />

dH16363<br />

<strong>CBS</strong>114399<br />

<strong>CBS</strong>117905<br />

<strong>CBS</strong>16654<br />

<strong>CBS</strong>114396<br />

<strong>CBS</strong>117904<br />

<strong>CBS</strong>114402<br />

<strong>CBS</strong>114397<br />

<strong>CBS</strong>85796<br />

<strong>CBS</strong>114394<br />

<strong>CBS</strong>86396<br />

<strong>CBS</strong>86196<br />

<strong>CBS</strong>16354<br />

<strong>CBS</strong>117896<br />

<strong>CBS</strong>85896<br />

<strong>CBS</strong>36270<br />

<strong>CBS</strong>26083<br />

<strong>CBS</strong>40696<br />

<strong>CBS</strong>100434<br />

<strong>CBS</strong>16054<br />

<strong>CBS</strong>114407<br />

<strong>CBS</strong>114405<br />

<strong>CBS</strong>114406<br />

dH14614<br />

K=5 K=4<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

unknown<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

unknown<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

Australia<br />

Venezuela<br />

Venezuela<br />

Mozambique<br />

Ug<strong>and</strong>a<br />

Australia<br />

Madagascar<br />

Australia<br />

Venezuela<br />

Venezuela<br />

Venezuela<br />

I<br />

II<br />

III<br />

Fig. 3. Phylogenetic tree (Neighbour-joining) of the C. carrionii complex based on EF1 (with grouping I–III) using the same strains of Fig. 1, generated using the HKY+G model.<br />

<strong>The</strong> model was calculated using ML in MrAic software. Bootstrap cut-off = 80 %. <strong>CBS</strong> 834.96 was taken as outgroup. Columns were generated with Structure software<br />

hypothesising K = 4 <strong>and</strong> K = 5, <strong>and</strong> alleles independent. Geographical origins in black refer to isolates from humans (chromoblastomycosis); origins in green refer to isolates<br />

from plant material.<br />

inoculation, <strong>and</strong> from the thorns directly adjacent to this area.<br />

Samples were rinsed with 4 % sodium hypochlorite for 3 min, <strong>and</strong><br />

subsequently washed in sterile distilled water for re-isolations, <strong>and</strong><br />

for histological study by means of light microscopy (Fernández-<br />

Zeppenfeldt et al. 1994).<br />

Experiments with spines of S. griseus<br />

Ninety cactus spines of 2.5 cm av. length were collected from<br />

a single S. griseus plant located near the home of the patient<br />

infected with <strong>CBS</strong> 114402, at approx. 2.5 m height, superficially<br />

sterilised, <strong>and</strong> divided into three groups, of which 30 spines were<br />

inoculated with <strong>CBS</strong> 114402 (C. carrionii), 30 with <strong>CBS</strong> 114405<br />

(Cladophialophora sp.) <strong>and</strong> 30 to be used as control, inoculated<br />

with a saline solution (Fig. 1D). A <strong>similar</strong> series composed of 90<br />

spines of 1.5 cm average length was collected at approx. 1 m<br />

height. All spines were incubated in sterile Petri dishes with filter<br />

paper (Whatman #1) with 2 mL saline solution; subsequently 0.1<br />

mL fungal suspension was applied. Twenty spines were analysed<br />

weekly by means of longitudinal sectioning with a h<strong>and</strong>-held<br />

microtome, cultured as above <strong>and</strong> observed microscopically until<br />

day 75 post incubation.<br />

www.studiesinmycology.org<br />

225


De Hoog et al.<br />

ROOT<br />

<strong>CBS</strong>114398<br />

<strong>CBS</strong>117892<br />

<strong>CBS</strong>16654<br />

<strong>CBS</strong>86196<br />

<strong>CBS</strong>41096<br />

<strong>CBS</strong>114404<br />

<strong>CBS</strong>114395<br />

<strong>CBS</strong>114392<br />

<strong>CBS</strong>85796<br />

<strong>CBS</strong>117903<br />

<strong>CBS</strong>117899<br />

<strong>CBS</strong>117909<br />

<strong>CBS</strong>114403<br />

<strong>CBS</strong>114399<br />

5<br />

<strong>CBS</strong>114397<br />

<strong>CBS</strong>85896<br />

<strong>CBS</strong>114402<br />

<strong>CBS</strong>114396<br />

<strong>CBS</strong>117896<br />

<strong>CBS</strong>117904<br />

dH16363<br />

<strong>CBS</strong>16354<br />

<strong>CBS</strong>86396<br />

<strong>CBS</strong>117906<br />

<strong>CBS</strong>117900<br />

<strong>CBS</strong>117891<br />

<strong>CBS</strong>117905<br />

<strong>CBS</strong>114394<br />

<strong>CBS</strong>117901<br />

<strong>CBS</strong>114393<br />

<strong>CBS</strong>40696<br />

<strong>CBS</strong>16054<br />

<strong>CBS</strong>100434<br />

<strong>CBS</strong>36270<br />

1 7<br />

4 6<br />

3 2 9<br />

10<br />

8<br />

<strong>CBS</strong>26083<br />

A<br />

B<br />

<strong>CBS</strong>114406<br />

<strong>CBS</strong>114405<br />

<strong>CBS</strong>114407<br />

A<br />

Unknown<br />

Australia<br />

Australia<br />

Australia<br />

Madagascar<br />

Mozambique<br />

D<br />

Ug<strong>and</strong>a<br />

C. carrionii<br />

C. yegresii<br />

Fig. 4. Reticulogram of South American strains of Cladophialophora species <strong>and</strong> strains from other continents (mentioned) constructed with T-rex software. ITS rDNA (with<br />

grouping A, B, D) was used as species tree <strong>and</strong> compared with the ß-tubulin gene tree. First 10 reticulations are shown with numbers.<br />

Statistics<br />

Survival of the cactus seedlings <strong>and</strong> collected plants following<br />

inoculation were evaluated using the X 2 -test (P = 0.05 was<br />

considered significant). Stem lesions resulting from inoculations<br />

were analysed with Student’s T-test (P = 0.01 was considered<br />

significant).<br />

RESULTS<br />

<strong>The</strong> rDNA ITS region was sequenced for 43 strains identified as<br />

C. carrionii based on morphology. Sequences of 16 additional<br />

strains were downloaded from GenBank. Five distantly related,<br />

unidentified cladophialophora-like species were added, with <strong>CBS</strong><br />

834.96 as outgroup. In C. carrionii, 203 positions were compared<br />

in ITS1, 158 in the 5.8S rRNA gene <strong>and</strong> 182 in ITS2 (Table 3).<br />

<strong>The</strong> sequences could be aligned with confidence over their entire<br />

lengths. Over the data set, 35 positions were polymorphic, of which<br />

33 were phylogenetically informative (Table 3), the two remaining<br />

being variable T-repeats near the ends of ITS1 <strong>and</strong> 2.<br />

For ITS sequences the AICc selected the TrN+G model (TrNG;<br />

Tamura & Nei 1993). <strong>The</strong> base frequency of ITS: T = 0.2467, C<br />

= 0.2897, A = 0.2247, G = 0.2390, TC = 0.5364, AG = 0.4636.<br />

<strong>The</strong> EF1 tree was built with substitution model HKY+G; the base<br />

frequency of EF1: T = 0.2990, C = 0.2665, A = 0.2123, G = 0.2221,<br />

TC = 0.5655, AG = 0.4345. <strong>The</strong> best model for BT2 sequences<br />

was the SYM+I+G (symmetrical model). <strong>The</strong> base frequency for<br />

BT2: T = 0.2255, C = 0.2953, A = 0.2463, G = 0.2328, TC = 0.5208,<br />

AG = 0.4792. Bootstrap values of the EF1 tree were calculated<br />

with PAUP using parsimony <strong>and</strong> with maxtrees set to 500 <strong>and</strong> 500<br />

replicates (data not shown). Total number of characters was 191 of<br />

which 101 were parsimony-informative. Tree length was 365 <strong>and</strong><br />

had the following indices: Consistency Index = 0.685, Retention<br />

Index = 0.542 <strong>and</strong> Homoplasy Index = 0.315.<br />

<strong>The</strong> original tree length, L o<br />

was 1 055, the tree length of the<br />

combined data, L c<br />

was 1 062. <strong>The</strong> resulting incongruence length<br />

difference L = (L c<br />

–L o<br />

) was 7 (P = 0.24). <strong>The</strong> observed ILD was not<br />

significantly greater than expected by chance <strong>and</strong> it was concluded<br />

that the sequences were congruent <strong>and</strong> could be used together in<br />

a combined analyses.<br />

Split decomposition based on the same alignment generated<br />

extensive recombination. <strong>The</strong> structure found with three loci was<br />

robust, with the exception of separation of <strong>CBS</strong> 834.96 <strong>and</strong> <strong>CBS</strong><br />

102227 with EF1 (Fig. 2).<br />

<strong>The</strong> core of the network, comprising the strains listed in Table 1,<br />

was analysed in more detail. With ITS, four groups were recognised<br />

(A–D; Table 1). (A) was the main group with 36 strains / sequences;<br />

FMC 248 differed only by a small T-repeat <strong>and</strong> was regarded as a<br />

member of (A). <strong>The</strong> remaining groups were smaller, differing from<br />

group (A) maximally by two consistent positions (Table 3). Group<br />

(C) mainly comprised sequences from GenBank <strong>and</strong> all originated<br />

from Abliz et al. (2004). One of the strains of group (C), IFM 4808,<br />

concerned a subculture of <strong>CBS</strong> 160.54, which is an original isolate<br />

of Trejos (1954) representing C. carrionii. Re-sequencing indicated<br />

that it was a member of group (B). Analysis of our electropherograms<br />

of this isolate was not suggestive of heterothallism. None of the<br />

positions characterising groups (A)–(C) were also found to differ in<br />

group (D), which deviated in 16 mutations in ITS1 <strong>and</strong> 8 in ITS2;<br />

17 of the mutations were transitions, 7 were transversions <strong>and</strong> 7<br />

indels. Group (D) was clearly distinct from the complex of (A)–(C),<br />

226


Cladophialophora carrionii complex<br />

Table 3. Nucleotide variability of ITS1-2 ribosomal DNA regions of<br />

Cladophialophora carrionii (A – C) <strong>and</strong> C. yegresii (D).<br />

rDNA domains (length), with variable nucleotide positions.<br />

ITS1 (201-203) A B C D<br />

16 C C C T<br />

17 T C T T<br />

19 T T T C<br />

51 A A A G<br />

57 A A A T<br />

90-92 TG- TG- TG- CGT<br />

101 T T T C<br />

103 C C C T<br />

104 G A G G<br />

106 A A A G<br />

114 T T T C<br />

122 T T T C<br />

132 C C C T<br />

137 A A A C<br />

141 C C C T<br />

145 - - - A<br />

163-170 6-10T 6-10T 6-10T TTGTATCT<br />

180 - - - A<br />

183 G G G A<br />

190 T/A A A A<br />

5.8S (158) Monomorphic<br />

ITS2 (178-182) A B C D<br />

36 C T T T<br />

48 T T G T<br />

49 T T T C<br />

51 - - - C<br />

114 C C C G<br />

140 A A A G<br />

155 - - - T<br />

178-179 -- -- -- CT<br />

with a total of 27 mutations.<br />

For multilocus analysis with ITS, EF1 <strong>and</strong> BT2 a smaller set<br />

of strains was compared. Sequences of the 205 bp long element<br />

of EF1 contained 32 phylogenetically informative mutations. Three<br />

entities were distinguished (I–III; Fig. 3). With BT2, three groups<br />

with the same composition were recognised. Strains of ITS group<br />

(C) were not available for study.<br />

On the basis of multilocus screening in BioNumerics, concordant<br />

groups (A)–(D) were tested with the Structure programme. When<br />

K was set at 4 or 5, consistent groupings were noted, indicated as<br />

I, II <strong>and</strong> III (Fig. 3), corresponding with ITS groups (A), (B) <strong>and</strong> (D),<br />

respectively in Table 1.<br />

<strong>The</strong> possibility that group (D) / (III) included a member of another,<br />

morphologically <strong>similar</strong> but phylogenetically unrelated group of fungi<br />

was excluded by SSU sequencing. Genera morphologically <strong>similar</strong><br />

to Cladophialophora, such as <strong>Cladosporium</strong> Link, Devriesia Seifert<br />

www.studiesinmycology.org<br />

& N.L. Nick., Phaeoramularia Munt.-Cvetk., Pseudocladosporium<br />

U. Braun <strong>and</strong> Stenella Syd. proved to be remote (data not shown).<br />

With T-rex, interaction between groups (B) <strong>and</strong> (D) was noted,<br />

rather than between groups (B) <strong>and</strong> (A), despite the high sequence<br />

<strong>similar</strong>ity of (A) <strong>and</strong> (B) (Fig. 4).<br />

Morphological observation revealed that representatives of<br />

ITS groups (A)–(C) generally had conidiophores that arise at right<br />

angles from creeping hyphae (Fig. 5), while those of (D) tend to<br />

be ascending, hyphae gradually becoming conidiophore-like. Since<br />

slight correspondence was found in independent markers <strong>and</strong><br />

phenetic criteria, we considered group (D) to represent a separate<br />

species, which is described as follows.<br />

Cladophialophora yegresii de Hoog, sp. nov. MycoBank<br />

MB500208. Figs 6, 7D–F.<br />

Etymology: Named after Francisco Yegres, Venezuelean<br />

mycologist.<br />

Coloniae in agaro PDA dicto 22 °C planae, olivaceo-virides, pulverulentae vel<br />

velutinae, margine integra; reversum olivaceo-atrum. Hyphae fertiles dilute olivaceovirides,<br />

ascendentes, paulatim in catenas conidiorum concolorium vertentes.<br />

Conidiorum catenae ramosae, conidia dilute olivaceo-viridia, levia et tenuitunicata,<br />

4.5–6 × 2.5 µm, faciliter liberata, cicatricibus modice pigmentatis. Chlamydosporae<br />

et cellulae zymosae absentes. Synanamorphe phialidica non visa. Teleomorphe<br />

ignota.<br />

Holotypus cultura sicca <strong>CBS</strong> H-18464 in herbarium <strong>CBS</strong> praeservatur.<br />

Colonies on PDA at 22 °C evenly olivaceous green, powdery to<br />

velvety, with entire margin; reverse olivaceous black. Fertile hyphae<br />

pale olivaceous green, ascending, gradually changing over into<br />

concolorous chains of conidia. Conidial system profusely branched.<br />

Conidia pale olivaceous green, smooth- <strong>and</strong> thin-walled, 4.5–6<br />

× 2.5 µm, detached rather easily, with slightly pigmented scars.<br />

Chlamydospores <strong>and</strong> yeast cells absent. Phialidic synanamorph<br />

not observed. Teleomorph unknown.<br />

Specimen examined: Venezuela, Falcon state, from asymptomatic Stenocereus<br />

griseus cactus, G. Fernández-Zeppenfeldt, <strong>CBS</strong> H-18464 holotype, culture ex-type<br />

<strong>CBS</strong> 114405 = UNEFM SgSR3.<br />

Notes: Of the 48 <strong>dematiaceous</strong> isolates obtained from 36 fragments<br />

of the cactus Stenocereus griseus, four strains originating from<br />

four different plants of S. griseus presented morphological <strong>and</strong><br />

physiological key characteristics of Cladophialophora carrionii or<br />

C. yegresii (de Hoog et al. 2000, 2006). Gelatin liquefaction was<br />

negative in all strains <strong>and</strong> the maximum growth temperature was<br />

37 °C. After identification to species level using sequence data (de<br />

Hoog et al. 2006), both C. carrionii <strong>and</strong> C. yegresii appeared to be<br />

among the strains isolated.<br />

A total of 256 plants obtained at the end of 1 yr from germlings,<br />

had ribs, spines, <strong>and</strong> an average height of 15 cm. <strong>The</strong> 96<br />

germlings inoculated with fungal suspensions of the test strains<br />

<strong>CBS</strong> 114402 (C. carrionii, clinical) <strong>and</strong> <strong>CBS</strong> 114405 (C. yegresii,<br />

environmental) remained without visible external lesions during the<br />

year of experimentation. Histological sections of the 96 inoculated<br />

plants consistently revealed internal growth of the fungi in their<br />

filamentous form. Muriform cells were not observed, neither on<br />

the epidermis, nor in the internal tissue, spines or roots. <strong>The</strong> reisolated<br />

cultures demonstrated the viability of the fungi during the<br />

entire experimental process: <strong>CBS</strong> 114402 (C. carrionii) was grown<br />

from 26 (54.16 %) of the plants <strong>and</strong> <strong>CBS</strong> 114405 (C. yegresii) in<br />

23 (47.90 %) of the plants. <strong>The</strong> X 2 test did not reveal significant<br />

differences between the isolates (X 2 c = 0.0729 < X 2 t = 3.84). <strong>The</strong><br />

32 control plants remained without external lesions, <strong>and</strong> in the<br />

227


De Hoog et al.<br />

Fig. 5. Microscopic morphology of C. carrionii, strain <strong>CBS</strong> 160.54. Conidiophore erect, i.e. mostly arising at 90° from creeping hypha (sketch). Scale bar = 10 µm.<br />

histological sections no internal or external fungal elements were<br />

observed. None of the fungi isolated from the control plants proved<br />

to be a species of Cladophialophora.<br />

With 96 plants with superficial application of spore suspension<br />

(48 plants for each isolate, either clinical or environmental) neither<br />

internal nor external lesions were observed. Histological sectioning<br />

did not reveal fungal elements in or on plant tissue. Short hyphal<br />

elements <strong>and</strong> meristematic cells were occasionally seen around<br />

<strong>and</strong> inside the outer layers of the spines. <strong>The</strong> re-isolated strains<br />

proved that the fungi survived during the entire experimental<br />

procedure: <strong>CBS</strong> 114402 (C. carrionii) was isolated from 32 (66.67<br />

%) plants <strong>and</strong> <strong>CBS</strong> 114405 (C. yegresii) from 33 (68.75 %). <strong>The</strong> X 2<br />

test did not detect significant differences in survival rates among<br />

the isolates (X 2 c = 0.4375 < X 2 t = 3.84).<br />

Mature plants inoculated using colonised toothpicks showed<br />

average scarring of 1.88 cm diam with C. carrionii <strong>and</strong> 1.33 cm<br />

diam with C. yegresii, around the point of inoculation. In histological<br />

sections of 100 plants, dark, septate hyphae with inflated elements<br />

were observed at the points of inoculation. Muriform cells were not<br />

observed. Re-isolated strains were evidence of isolate viability:<br />

228


Cladophialophora carrionii complex<br />

Fig. 6. Microscopic morphology of C. yegresii, strain <strong>CBS</strong> 114405. Conidiophore ascending, i.e. mostly emerging from hyphal end that is gradually growing upwards to become<br />

a conidiophore (sketch). Scale bar = 10 µm.<br />

<strong>CBS</strong> 114402 (C. carrionii) was grown from 36 (72 %) plants <strong>and</strong> <strong>CBS</strong><br />

114405 (C. yegresii) from 30 (60 %). <strong>The</strong> fungi could not be isolated<br />

from spines. <strong>The</strong> 50 plants used as controls showed scarring of<br />

1.06 cm diam on average around the point of inoculation. No fungal<br />

elements were seen in direct examinations <strong>and</strong> histological sections<br />

of these plants. <strong>The</strong> retro-cultures were negative. <strong>The</strong> scarring<br />

responses of the plants to the clinical strain, environmental strain<br />

<strong>and</strong> control proved to be highly significant:<br />

Clinical <strong>CBS</strong> 114402 vs. environmental <strong>CBS</strong> 114405: þ = 0.000832,<br />

P = 0.01;<br />

Clinical <strong>CBS</strong> 114402 vs. control: þ = 0.00003128, P = 0.01;<br />

Environmental <strong>CBS</strong> 114405 vs. control: þ = 0.005343, P = 0.01.<br />

Spines 2.5 cm av. in length, seeded with suspensions of <strong>CBS</strong><br />

114402 (C. carrionii) <strong>and</strong> <strong>CBS</strong> 114405 (C. yegresii), developed<br />

toruloid hyphal elements with some dark, swollen cells <strong>similar</strong><br />

to muriform cells known in human tissue. <strong>The</strong> re-isolated strains<br />

proved the species to survive during the experimental procedure<br />

(< 75 d). Similar results were obtained with the spines 1.5 cm av.<br />

in length. No cladophialophora-like fungi were isolated from the<br />

controls.<br />

www.studiesinmycology.org<br />

229


De Hoog et al.<br />

Fig. 7. Conidial morphology in selected branches of (upper row: A–C) C. carrionii, strain <strong>CBS</strong> 260.83; (lower row: D–F) C. yegresii, strain <strong>CBS</strong> 114405. In this respect the two<br />

species are identical. Scale bar = 10 µm.<br />

DISCUSSION<br />

Taxonomy of Cladophialophora<br />

Judging from SSU rDNA phylogeny data, all Cladophialophora<br />

species that are consistently associated with pathology to humans<br />

belong to the Herpotrichiellaceae in the order Chaetothyriales<br />

(Haase et al. 1999). Within this order, the <strong>genus</strong> Cladophialophora<br />

is polyphyletic. Conidia of all species are produced in branched<br />

chains on poorly differentiated hyphae. This very simply structured<br />

conidial system may lead to confusion with morphologically<br />

<strong>similar</strong> but unrelated fungi that are encountered as contaminants<br />

in the hospital environment. <strong>The</strong> <strong>genus</strong> <strong>Cladosporium</strong> comprises<br />

ubiquitous airborne fungi which mostly have erect, more or less<br />

differentiated conidiophores, <strong>and</strong> dark conidial scars. <strong>The</strong>y are<br />

associated with Davidiella Crous & U. Braun teleomorphs <strong>and</strong><br />

belong to the Dothideomycetes, family Davidiellaceae (Braun et al.<br />

2003, Schoch et al. 2006). Pseudocladosporium was introduced by<br />

Braun (1998) with three species differing from Cladophialophora<br />

mainly by intercalary hyphal cells with lateral extensions that bear<br />

conidial chains, having Caproventuria U. Braun teleomorphs (see<br />

Crous et al. 2007 – this volume). <strong>The</strong> group is classified in the<br />

Venturiaceae <strong>and</strong> Mycosphaerellaceae in the Dothideales (Braun et<br />

al. 2003). <strong>The</strong> anamorph <strong>genus</strong> Devriesia comprises thermophilic<br />

saprobes with a cladophialophora-like appearance <strong>and</strong> producing<br />

230


Cladophialophora carrionii complex<br />

dark, multi-celled chlamydospores alongside the hyphae.<br />

Phylogenetically this <strong>genus</strong> is related to the Mycosphaerellaceae,<br />

in the Dothideomycetes (Seifert et al. 2004).<br />

Cladophialophora carrionii was originally introduced by Trejos<br />

(1954) on the basis of 46 strains from Venezuela, Australia <strong>and</strong><br />

South Africa. He did not indicate a holotype. For this reason isolate<br />

Trejos 27 = Emmons 8619 = <strong>CBS</strong> 160.54, the first strain mentioned<br />

by Trejos (1954), is selected here as representative for C. carrionii.<br />

A dried specimen of this strain has been deposited as lectotype<br />

in the Herbarium of the Centraalbureau voor Schimmelcultures as<br />

<strong>CBS</strong> H-18465.<br />

<strong>The</strong> ex-type strain of Cladophialophora ajelloi Borelli, <strong>CBS</strong><br />

260.83, proved to be indistinguishable from C. carrionii, which was<br />

also known to be able to produce phialides in addition to catenate<br />

conidia (Honbo et al. 1984). Remarkably, a strain identified as C.<br />

ajelloi from Samoa (<strong>CBS</strong> 259.83; Goh et al. 1982) proved to be<br />

related to but consistently different from all strains of the C. carrionii<br />

complex. <strong>The</strong> 43-year-old male patient in otherwise good health<br />

carrying this fungus had a 5 × 3 cm erythematous, scaling lesion<br />

on his arm. Muriform cells were present in superficial dermis <strong>and</strong><br />

stratum corneum. This clearly represents yet another agent of<br />

human chromoblastomycosis. <strong>The</strong> name C. ajelloi is not available<br />

for this taxon, as this is a synonym of C. carrionii. <strong>The</strong> taxon will be<br />

formally described in a forthcoming paper.<br />

Members of ITS groups (A)–(D) were shown to be close to<br />

each other in SSU phylogeny (data not shown) underlining that all<br />

analysed species were correctly assigned to Cladophialophora.<br />

This <strong>genus</strong> was defined by melanised acropetal chains of conidia,<br />

near absence of conidiophores, <strong>and</strong> phylogenetic affinity to the<br />

order Chaetothyriales. Strains (A)–(D) clustered in a clade which<br />

contained a mixture of species of Cladophialophora, Fonsecaea<br />

Negroni <strong>and</strong> Phialophora Medlar. From a point of view of human<br />

disease, the species of the clade were known as agents of brain<br />

infection [C. bantiana (Sacc.) de Hoog et al., F. monophora (M. Moore<br />

& F.P. Almeida) de Hoog et al.], disseminated disease [C. devriesii<br />

(A.A. Padhye & Ajello) de Hoog et al.], cutaneous disease [C. boppii<br />

(Borelli) de Hoog et al.] <strong>and</strong> particularly chromoblastomycosis (C.<br />

carrionii, Fonsecaea, Phialophora).<br />

Diversity of Cladophialophora carrionii / C. yegresii<br />

Infraspecific variability was observed within C. carrionii. <strong>The</strong> groups<br />

(A)–(C) were separated on the basis of five mutations in the ITS<br />

region, which were supported by mutations in EF1 <strong>and</strong> BT2, as<br />

confirmed by analysis in Structure, where the same separation<br />

(K = 5) of entities was observed. Furthermore, K = 4 unites groups<br />

(B/II) <strong>and</strong> (D/III), despite the fact that the sequence of (B) is more<br />

close to those of (A). With T-rex software a <strong>similar</strong> relationship<br />

between [(B), C. carrionii] <strong>and</strong> [(D), C. yegresii] was noted,<br />

suggesting horizontal gene flow between these entities. This is<br />

remarkable, since (C) strains predominantly inhabit remote deserts<br />

in Madagascar <strong>and</strong> Australia, while (D) is found in equally remote<br />

localities in Venezuela. Extensive reticulation was observed in all<br />

genes with SplitsTree. With ITS <strong>and</strong> BT2, <strong>CBS</strong> 834.96 <strong>and</strong> <strong>CBS</strong><br />

102227 cluster closely together, while in the more variable EF1<br />

data these are all widely apart, suggesting that in Cladophialophora<br />

other mechanisms than recombination may occur.<br />

Group (C) contained ITS sequences taken from the public<br />

domain, originating from a single study (Abliz et al. 2004).<br />

Remarkably, strain IFM 4808 found in group (C) on the basis of data<br />

from Abliz et al. (2004), was the same isolate as <strong>CBS</strong> 160.54, which<br />

was found repeatedly in group (B) in our data set (Table 1). A <strong>similar</strong><br />

www.studiesinmycology.org<br />

phenomenon was observed with strain IFM 41444 = <strong>CBS</strong> 863.96,<br />

of which GenBank deposition AB109169 consistently deviated from<br />

our data in a frequently observed mutation. A possible explanation<br />

of these consistent sequence conflicts is heterozygosity. Although<br />

most chaetothyrialean fungi are supposed to be haploid (Szaniszlo<br />

2002; Zeng et al. 2007), some strains have a double DNA content<br />

in yeast cells (Ohkusu et al. 1999). Teleomorphs are not known<br />

in Cladophialophora <strong>and</strong> related black yeasts, but many species<br />

are known to form profuse hyphal anastomoses (de Hoog et al.<br />

2006), allowing parasexual processes <strong>and</strong> mitotic recombination.<br />

However, all electropherograms including those from the study of<br />

Abliz et al. (2004), which were kindly sent by K. Fukushima (Chiba,<br />

Japan), were unambiguous, without double peaks. This matched<br />

with the observation of preponderant clonality despite frequent<br />

anastomoses in Exophiala J.W. Carmich. (Zeng et al. 2007). An<br />

alternative explanation might be the occurrence of paralogous ITS<br />

repeats, as reported earlier in Fusarium Link (O’Donnell & Cigelnik<br />

1997).<br />

<strong>The</strong> remaining diversity within C. carrionii as confirmed by<br />

Structure shows some geographical structuring of populations,<br />

in that group (A) does not occur in Asia, group (B) is limited to<br />

Australia <strong>and</strong> Africa, <strong>and</strong> group (C) has thus far only been reported<br />

from Asia. <strong>The</strong> wide distribution of most genotypes suggests,<br />

however, that worldwide occurrence is likely to become apparent<br />

when more strains have been analysed. All climate zones where<br />

C. carrionii was isolated were semi-arid to arid, desert-like.<br />

Genotypes were not limited to the endemic semi-arid areas, <strong>and</strong><br />

thus a relatively rapid vector of dispersal has to be hypothesised<br />

enabling the fungus to cross climate zones where the saprobic<br />

phase is unable to survive. Kawasaki et al. (1993) analysed three<br />

further loci in mtDNA using RFLP. Only some of their strains were<br />

available for sequencing. <strong>The</strong>se had all identical mtDNA profiles,<br />

with the exception of IFM 4808 = <strong>CBS</strong> 160.54, that differed in two<br />

markers (Table 1). If we assume that there is no real separation of<br />

ITS groups (A) <strong>and</strong> (C) (see above), the conclusion is warranted<br />

that mtDNA allows distinction of polymorphism at the same level of<br />

diversity as detected in this study with ITS, EF1 <strong>and</strong> BT2.<br />

South America harbours group (D) which represents a second<br />

species, C. yegresii. This species thus far has not been found<br />

on humans, <strong>and</strong> seems to be restricted to living Stenocereus<br />

cactus plants. Nishimura et al. (1989) published a strain from<br />

chromoblastomycosis in China which matched the morphology of<br />

strains now classified as C. yegresii, but as far as we are aware this<br />

strain has not been sequenced.<br />

Ecology <strong>and</strong> virulence of Cladophialophora carrionii / C.<br />

yegresii<br />

Cladophialophora carrionii was preponderantly found as an agent<br />

of human infection <strong>and</strong> only occasionally on dead plant debris,<br />

mainly seceded cactus needles. <strong>The</strong> only three strains available<br />

of C. yegresii were isolated from living, asymptomatic Stenocereus<br />

plants surrounding the cabin of a symptomatic patient from whom C.<br />

carrionii, <strong>CBS</strong> 114402 was isolated. Although in some publications<br />

convincing evidence was presented that infections originate from<br />

puncture by plant material (e.g., Salgado et al. 2004), it now<br />

becomes clear that the environmental look-alikes of clinical strains<br />

do not necessarily belong to the same species (Crous et al. 2006,<br />

Mostert et al. 2006), but may be members of other, related taxa with<br />

slightly different ecology; an unambiguous connection between a<br />

clinical <strong>and</strong> an environmental strain still has to be proven.<br />

<strong>The</strong> endemic area of the two species, C. carrionii <strong>and</strong> C.<br />

231


De Hoog et al.<br />

yegresii, has a semi-arid climate, with average yearly temperatures<br />

of 24 ºC, scarce rainfall (up to 600 annual mL) <strong>and</strong> is located at<br />

moderate altitude (up to 500 m) (Borelli 1979, Richard-Yegres &<br />

Yegres 1987). <strong>The</strong> l<strong>and</strong>scape is dominated by large cacti <strong>and</strong> other<br />

xerophytes. Stenocereus griseus is a columnar American cactus<br />

with a very strong, protective external epidermis that allows the<br />

accumulation of water in the shaft <strong>and</strong> enables tolerance of extreme<br />

drought. <strong>The</strong> species produces ovoidal, thorny fruits of about 5 cm<br />

diam, which are commonly eaten by the local population. It has<br />

therefore been suggested that patients with chromoblastomycosis<br />

acquire their infection by traumatic inoculation with cactus spines,<br />

<strong>similar</strong> to the supposed infection process of Madurella mycetomatis<br />

(Laveran) Brumpt in the arid climate of Africa (Ahmed et al. 2002).<br />

<strong>The</strong> frequent occurrence of 16 / 1 000 for chromoblastomycosis in<br />

areas endemic for Cladophialophora in Venezuela (Yegres et al.<br />

1985; Yegűez-Rodriguez et al. 1992) indicates a marked invasive<br />

potential for C. carrionii. Local goat-keepers are particularly at risk:<br />

in 1984, 14 of 18 patients investigated had these occupational<br />

characteristics (Yegres et al. 1985). Nevertheless, virulence of C.<br />

carrionii is low when inoculated into the footpads of mice (Yegres<br />

et al. 1998); also an environmental strain of C. carrionii failed to<br />

produce lesions in mice <strong>and</strong> in a volunteer (Richard-Yegres &<br />

Yegres 1987).<br />

We performed inoculation experiments with C. carrionii <strong>and</strong><br />

C. yegresii using freshly grown, healthy cacti in the greenhouse.<br />

<strong>The</strong> plants were followed over a 1-yr period; during all this time<br />

the control plants remained without lesions. Both Cladophialophora<br />

strains were able to produce infection when syringe-inoculated<br />

deep into young cactus tissue. Histopathology showed septate<br />

hyphae between host cells, <strong>and</strong> the shaft was maintained over<br />

prolonged periods without causing visible damage. This absence<br />

of appreciable destruction would categorise them as endophytes.<br />

Cactus tissue is rich in carbohydrates, vitamins <strong>and</strong> minerals (Vélez<br />

& Chávez 1980) which may promote endophyte growth.<br />

In contrast, suspensions applied superficially lead to growth<br />

on <strong>and</strong> in spines only. <strong>The</strong> absence of infection after superficial<br />

application indicates that the fungi are unable to invade healthy<br />

plant tissue from the surface <strong>and</strong> thus they cannot be characterised<br />

as obligatory phytopathogens.<br />

<strong>The</strong> two species differed in the degree of scarring after traumatic<br />

inoculation into mature plants: the clinical strain C. carrionii was<br />

consistently more virulent than C. yegresii that originated from<br />

the same host plant. Both species showed the same viability in<br />

re-isolated cultures. In nature, the fungi are likely to invade only<br />

when the integrity of the epidermis is broken, as happens e.g.<br />

by goat feeding or transmission by sap-sucking birds or piercing<br />

insects. <strong>The</strong>y also show the same transformation to meristematic<br />

morphology (González et al. 1990) when entering hard spine<br />

tissue. A possible trigger for this conversion is the dominance of<br />

lignin in the spines. Survival on <strong>and</strong> in spines is enhanced by their<br />

capturing of atmospheric water formed after nightly condensation.<br />

<strong>The</strong> fact that superficial application leads to colonisation around <strong>and</strong><br />

inside the spines suggests that the spines play a role in mechanic<br />

dispersion of the fungi.<br />

A possibly coincidental mechanism of dispersal might be<br />

traumatic inoculation into living tissue of humans or animals, where<br />

the same muriform cells are formed, defining the skin disease<br />

chromoblastomycosis. It may be questioned whether animal/<br />

human inoculation plays a role in the evolution of the fungus. ITS<br />

differences between the two species are observed in 23 positions,<br />

with a ratio of transitions : transversions of 2 : 1 (Table 3). Thus<br />

no saturation of mutations has taken place <strong>and</strong> the diversification<br />

can be regarded as an example of recent sympatric speciation.<br />

Cladophialophora carrionii is widely distributed, <strong>and</strong> shows a higher<br />

degree of diversity than C. yegresii. This would be suggestive for<br />

a longer evolutionary time span of existence <strong>and</strong> C. carrionii then<br />

should be regarded as ancestral to C. yegresii, with the latter<br />

showing a founder effect due to the absence of polymorphisms.<br />

However, such an order of event (a host jump from humans to<br />

cactus) is difficult to imagine. It is more likely that C. yegresii is<br />

the original cactus endophyte exhibiting extremotolerance via its<br />

muriform cells. T-rex data suggested a more direct connection of<br />

C. yegresii with African <strong>and</strong> Australian rather than Venezuelean<br />

strains of C. carrionii. We suppose that the low degree of observed<br />

variation in C. yegresii is not a founder effect, but rather a sampling<br />

effect, as living cacti have thus far not been studied outside the<br />

framework of our study on the patient with Cladophialophora<br />

chromoblastomycosis. <strong>The</strong> difference in virulence may be simply<br />

explained by C. carrionii, which lives as a saprobe on dead cactus<br />

debris for part of its life cycle, <strong>and</strong> is less adapted to an endophytic<br />

life style.<br />

Cladophialophora cf. carrionii is known to occur on lignified<br />

materials, such as wood chips of Eucalyptus crebra <strong>and</strong> wooden<br />

remains of Prosopis juliflora <strong>and</strong> Stenocereus griseus (Riddley<br />

1957, Yegres et al. 1985, Fernández-Zeppenfeldt et al. 1994). This<br />

does not exclude a certain degree of pathogenicity to humans, as<br />

also pathogens like Cryptococcus neoformans (Sanfelice) Vuill.<br />

are known to have an essential part of their life cycle in hollows<br />

of Eucalyptus trees. Cryptococcus neoformans produces diphenol<br />

oxidase to degrade lignin, an aromatic polymer in the cell wall of<br />

plants <strong>and</strong> a component of wood (Cabral 1999). Similar degradation<br />

pathways are present in Cladophialophora carrionii (Prenafeta-<br />

Boldú et al. 2006).<br />

<strong>The</strong> natural occurrence of C. carrionii <strong>and</strong> C. yegresii in<br />

association with xerophytes has been proven, but their environmental<br />

route of dispersal is still unknown. As transformation to meristematic<br />

cells takes place when the hyphae reach the spines <strong>and</strong> on dead<br />

spines, the muriform cell apparently is the extremotolerant phase<br />

of the species. <strong>The</strong> conidial anamorph can be found sporulating on<br />

rotten spines directly after rainfall (Richard-Yegres & Yegres 1987),<br />

but as the fungus has thus far never been isolated from outside air,<br />

it is still unclear how a new host plant is reached.<br />

<strong>The</strong> behaviour of C. carrionii on humans, provoking the very<br />

characteristic disease, chromoblastomycosis, of which the agents<br />

are limited to the ascomycete family Herpotrichiellaceae (de Hoog<br />

et al. 2000) is puzzling. In humans, the extremophilic muriform<br />

anamorph is expressed rather than hyphae, <strong>and</strong> thus humans do<br />

not seem a natural reservoir of the fungus. Nevertheless some<br />

acquired cellular immunity seems to be involved. Albornoz et al.<br />

(1982) demonstrated that a significant share of the local population<br />

of goat keepers (Yegres et al. 1985) is asymptomatically infected<br />

with C. carrionii; Iwatsu et al. (1982) detected cutaneous delayed<br />

hypersensitivity in rats experimentally-infected with agents of<br />

chromoblastomycosis. With murine experimental infection of<br />

the related fungus Fonsecaea pedrosoi, Ahrens et al. (1989)<br />

found enlargement <strong>and</strong> metastasis of lesions in athymic but not<br />

in normal mice, or in mice with defective macrophage function.<br />

Several authors (Kurita 1979, Nishimura & Miyaji 1981, Polak<br />

1984) observed a significant role of acquired cellular immunity in F.<br />

pedrosoi, while Cardona-Castro & Agudelo-Flórez (1999) obtained<br />

chronic infection in immunocompetent mice when inoculated<br />

intraperitoneally. Garcia Pires et al. (2002) noted an unbalance<br />

between protective Th1 <strong>and</strong> less efficient Th2 responses. <strong>The</strong><br />

possible host response leads to different clinical types, referred<br />

232


Cladophialophora carrionii complex<br />

to as tuberculoid <strong>and</strong> suppurative granuloma, respectively.<br />

<strong>The</strong> existence of genetic constitutional factors in susceptibility is<br />

underlined by a marked frequency of family relationships among<br />

symptomatic individuals (Yegűez-Rodriguez et al. 1992). <strong>The</strong><br />

disease is not observed in local animals such as goats, possibly<br />

due to their high body temperature (≈ 39 °C). Nevertheless, hyphal<br />

fragments artificially inoculated into goats led to transformation into<br />

muriform cells, but the lesions disappeared within 60 d (Martínez<br />

et al. 2005). Further animal experiments using strains identified<br />

according to new taxonomy will be necessary to answer questions<br />

on the role of the fungus on warm-blooded animals.<br />

ACKNOWLEDGEMENTS<br />

<strong>The</strong> authors are indebted to F. Yegres, N. Richard-Yegres <strong>and</strong> V. Maigualida Pérez-<br />

Blanco for providing a large set of strains for study <strong>and</strong> information on strain ecology,<br />

<strong>and</strong> to R.C. Summerbell for helpful suggestions on the manuscript. P. Abliz <strong>and</strong> K.<br />

Fukushima are acknowledged for making sequence data available for study. K.F.<br />

Luijsterburg is thanked for technical assistance. <strong>The</strong> study was supported in part<br />

by the Fondo Nacional de Investigaciones Cientificas y Tecnológicas (Fonacit),<br />

Venezuela.<br />

REFERENCES<br />

Abliz P, Fukushima K, Takizawa K, Nishimura K (2004). Specific oligonucleotide<br />

primers for identification of Cladophialophora carrionii, a causative agent of<br />

chromoblastomycosis. Journal of Clinical Microbiology 42: 404–407.<br />

Ahmed AOA, Adelmann D, Fahal A, Verbrugh H, Belkum A van, Hoog GS de (2002).<br />

Environmental occurrence of Madurella mycetomatis, the major agent of human<br />

eumycetoma in Sudan. Journal of Clinical Microbiology 40: 1031–1036.<br />

Ahrens J, Graybill JR, Abishawl A, Tio FO, Rinaldi MG (1989). Experimental murine<br />

chromomycosis mimicking chronic progressive human disease. American<br />

Journal of Tropical Medicine <strong>and</strong> Hygiene 40: 651–658.<br />

Albornoz MB, Marin C de, Iwatsu T (1982). Estudio epidemiologico de un area<br />

endemica para cromomicosis en el Estado Falcon. Investigaciόn Clínica 23:<br />

219–228.<br />

B<strong>and</strong>elt HJ, Dress AW (1992). Split decomposition: a new <strong>and</strong> useful approach to<br />

phylogenetic analysis of distance data. Molecular Phylogenetics <strong>and</strong> Evolution<br />

1: 242–252.<br />

Borelli D (1972). Significado del dimorfismo de ciertos hongos parásitos.<br />

Mycopathologia et Mycologia Applicata 46: 237–239.<br />

Borelli D (1979). Reservarea de algunos agentes de micosis. Medicina Cutánea<br />

4: 367–370.<br />

Braun U (1998). A monograph of Cercosporella, Ramularia <strong>and</strong> allied genera<br />

(phytopathogenic hyphomycetes), Vol. 2. IHW-Verlag, Eching.<br />

Braun U, Crous PW, Dugan F, Groenewald JZ, Hoog GS de (2003). Phylogeny <strong>and</strong><br />

taxonomy of <strong>Cladosporium</strong>-like hyphomycetes, including Davidiella gen. nov.,<br />

the teleomorph of <strong>Cladosporium</strong> s. str. Mycological Progress 2: 3–18.<br />

Burnham KP, Anderson DR (2002). Model selection <strong>and</strong> multimodel inference, a<br />

practical information-theoretic approach. 2nd ed, Springer, New York.<br />

Cabral L (1999). Wood, animals <strong>and</strong> human beings as reservoirs for human<br />

Cryptococcus neoformans infection. Revista Iberoamericana de Micologia 16:<br />

77–81.<br />

Carbone I, Kohn LM (1999). A method for designing primer sets for speciation<br />

studies in filamentous ascomycetes. Mycologia 91: 553–556.<br />

Cardona-Castro N, Agudelo-Flórez P (1999). Development of a chronic<br />

chromoblastomycosis model in immunocompetent mice. Medical Mycology 37:<br />

81–83.<br />

Cermeño J, Torres J (1998). Método espectrofotométrico en la preparación del<br />

inóculo de hongos dematiaceos. Revista Iberoamericana de Micologia 15:<br />

155–157.<br />

Clausnitzer I (1978). Germinación de las semillas del dato o frutos del cardón,<br />

Lemaireocereus griseus (Haw) Britt & Rose. Revista de la Facultad de<br />

Agronomia de la Universidad del Zulia 5: 351–365.<br />

Crous PW, Schubert K, Braun U, Hoog GS de, Hocking AD, Shin H-D, Groenewald<br />

JZ (2007). Opportunistic, human-pathogenic species in the Herpotrichiellaceae<br />

are phenotypically <strong>similar</strong> to saprobic or phytopathogenic species in the<br />

Venturiaceae. Studies in Mycology 58: 185–217.<br />

Crous PW, Slippers B, Wingfield MJ, Rheeder J, Marasas WFO, Philips AJL, Alves<br />

A, Burgess T, Barber P, Groenewald JZ (2006). Phylogenetic lineages in the<br />

Botryosphaeriaceae. Studies in Mycology 55: 235–253.<br />

Falush D, Stephens M, Pritchard JK (2003). Inference of population structure:<br />

Extensions to linked loci <strong>and</strong> correlated allele frequencies. Genetics 164:<br />

1567–1587.<br />

Felsenstein J (1993). PHYLIP (phylogeny inference package), version 3.6a2.<br />

Department of Genetics, Univ. Washington, Seattle.<br />

Fernández-Zeppenfeldt G, Richard-Yegres N, Yegres F, Hernández R (1994).<br />

<strong>Cladosporium</strong> carrionii: hongo dimórfico en cactáceas de la zona endémica<br />

para la cromomicosis en Venezuela. Revista Iberoamericana de Micologia 11:<br />

61–63.<br />

Garcia Pires d’Ávila SC, Pagliari C, Seixas Duarte MI (2002). <strong>The</strong> cell-mediated<br />

immune reaction in the cutaneous lesion of chromoblastomycosis <strong>and</strong> their<br />

correlation with different clinical forms of the disease. Mycopathologia 156:<br />

51–60.<br />

Glass NL, Donaldson G (1995). Development of primer sets designed for use with<br />

PCR to amplify conserved genes from filamentous ascomycetes. Applied <strong>and</strong><br />

Environmental Microbiology 61: 1323–1330.<br />

Goh KS, Padhye AA, Ajello L (1982). A Samoan case of chromoblastomycosis<br />

caused by Cladophialophora ajelloi. Sabouraudia 20: 1–5.<br />

González R, Caleiras E, Guanipa O, Garcia-Tamayo J, Siva L, Colina C, Fernández-<br />

Zeppenfeldt G, Pacheco I (1990). Chromomycosis: ultrastructural characteristics<br />

of the suppurative granuloma in <strong>Cladosporium</strong> carrionii skin lesions. Jornada<br />

Microscopica Electrónica 4: 133–134.<br />

Guindon S, Gascuel O (2003). A simple, fast, <strong>and</strong> accurate algorithm to estimate<br />

phyogenies by maximum likelihood. Systematic Biology 52: 696–704.<br />

Haase G, Sonntag L, Melzer-Krick B, Hoog GS de (1999). Phylogenetic interference<br />

by SSU-gene analysis of members of the Herpotrichiellaceae with special<br />

reference to human pathogenic species. Studies in Mycology 43: 80–97.<br />

Honbo S, Padhye AA, Ajello L (1984). <strong>The</strong> relationship of <strong>Cladosporium</strong> carrionii to<br />

Cladophialophora ajelloi. Sabouraudia 22: 209–218.<br />

Hoog GS de, Guarro J, Gené J, Figueras MJ (2000). Atlas of Clinical Fungi, 2 nd<br />

ed. Centraalbureau voor Schimmelcultures, Utrecht <strong>and</strong> Universitat Rovira i<br />

Virgili, Reus.<br />

Hoog GS de, Vicente VA, Caligiorne RB, Kantarcioglu AS, Tintelnot K, Gerrits<br />

van den Ende AHG, Haase G (2003). Species diversity <strong>and</strong> polymorphism in<br />

the Exophiala spinifera clade containing opportunistic black yeast-like fungi.<br />

Journal of Clinical Microbiology 41: 4767–4778.<br />

Hoog GS de, Zeng JS, Harrak MJ, Sutton DA (2006). Exophiala xenobiotica sp. nov.,<br />

an opportunistic black yeast inhabiting environments rich in hydrocarbons.<br />

Antonie van Leeuwenhoek 90: 257–268.<br />

Huson DH, Bryant D (2006). Application of phylogenetic networks in evolutionary<br />

studies. Molecular Biology <strong>and</strong> Evolution 23: 254–267.<br />

Iwatsu T, Miyaji M, Taguchi H, Okamoto S (1982). Evaluation of skin test for<br />

chromoblastomycosis using antigens prepared from culture filtrates of<br />

Fonsecaea pedrosoi, Phialophora verrucosa, Wangiella dermatitidis <strong>and</strong><br />

Exophiala jeanselmei. Mycopathologia 77: 59–64.<br />

Kawasaki M, Ishizaki H, Miyaji M, Nishimura K, Matsumoto T, Honbo S, Muir D<br />

(1993). Molecular epidemiology of <strong>Cladosporium</strong> carrionii. Mycopathologia<br />

124: 149–152.<br />

Kurita N (1979). Cell-mediated immune responses in mice infected with Fonsecaea<br />

pedrosoi. Mycopathologia 68: 9–15.<br />

Lavelle P (1980). Chromoblastomycosis in Mexico. Pan American Health<br />

Organization Scientific Publications 396: 235–247.<br />

Makarenkov V (2001). T-Rex: reconstructing <strong>and</strong> visualizing phylogenetic trees <strong>and</strong><br />

reticulation networks. Bioinformatics 17: 664–668.<br />

Makarenkov V, Legendre P (2004). From a phylogenetic tree to a reticulated<br />

network. Journal of Computational Biology 11: 195–212.<br />

Marques SG, Pedroso Silva CM, Saldanha PC, Rezende MA, Vicente VA, Queiros-<br />

Telles F, Lopes Costa JM (2006). Isolation of Fonsecaea pedrosoi from the shell<br />

of the babassu coconut (Orbignya phalerata Martius) in the Amazon region of<br />

Maranhão Brazil. Japanese Journal of Medical Mycology 47: 305–311.<br />

Martínez E, Rey Valieron C, Yergres F, Reyes R (2005). El caprino: aproximación<br />

a un modelo animal en la cromomicosis humana. Investigaciόn Clínica 46:<br />

131–138.<br />

Mendoza L, Karuppayil SM, Szaniszlo PJ (1993). Calcium regulates in vitro<br />

dimorphism in chromoblastomycotic fungi. Mycoses 36: 157–164.<br />

Mostert L, Groenewald JZ, Summerbell RC, Gams W, Crous PW (2006). Taxonomy<br />

<strong>and</strong> pathology of Togninia (Diaporthales) <strong>and</strong> its Phaeoacremonium anamorphs.<br />

Studies in Mycology 54: 1–115.<br />

Nishimura K, Miyaji M (1981). Defense mechanisms of mice against Fonsecaea<br />

pedrosoi infection. Mycopathologia 76: 155–166.<br />

Nishimura K, Miyaji M, Taguchi H, Wang DL, Li RY, Meng ZH (1989). An ecological<br />

study on pathogenic <strong>dematiaceous</strong> fungi from China. In: Current problems of<br />

opportunistic fungal infections. Proceedings of the 4 th International Symposium<br />

of the Research Center for Pathogenic Fungi <strong>and</strong> Microbial Toxicoses, Chiba<br />

University, Chiba: 17–20.<br />

Nyl<strong>and</strong>er JAA (2004). MrAic.pl. Programme distributed by the author. Evolutionary<br />

www.studiesinmycology.org<br />

233


De Hoog et al.<br />

Biology Centre, Uppsala University.<br />

O’Daly JA (1943). La cromoblastomicosis en Venezuela. Memorias Primera Jornada<br />

Venezolana de Venereología. Caracas.<br />

O’Donnell K, Cigelnik E (1997). Two divergent intragenomic rDNA ITS2 types within<br />

a monophyletic lineage of the fungus Fusarium are nonorthologous. Molecular<br />

<strong>and</strong> Phylogenetic Evolution 7: 103–116.<br />

Ohkusu M, Yamaguchi M, Hata K, Yoshida S, Tanaka S, Nishimura K, Hoog GS de,<br />

Takeo K (1999). Cellular <strong>and</strong> nuclear characteristics of Exophiala dermatitidis.<br />

Studies in Mycology 43: 143–150.<br />

Polak A (1984). Experimental infection of mice by Fonsecaea pedrosoi <strong>and</strong> Wangiella<br />

dermatitidis. Sabouraudia 22: 167–169.<br />

Posada D, Cr<strong>and</strong>all KA (1998). Modeltest: testing the model of DNA substitution.<br />

Bioinformatics 14: 817–818.<br />

Prenafeta-Boldú FX, Summerbell R, Hoog GS de (2006). Fungi growing on aromatic<br />

hydrocarbons: biotechnology’s unexpected encounter with biohazard? FEMS<br />

Microbiology Reviews 30: 109–130.<br />

Pritchard JK, Stephens M, Donnelly P (2000). Inference of population structure<br />

using multilocus genotype data. Genetics 155: 945–959.<br />

Richard-Yegres N, Yegres F (1987). <strong>Cladosporium</strong> carrionii en vegetacion xerofila:<br />

aislamiento en una zona endemica para la cromomicosis en Venezuela.<br />

Dermatologia Venezuelana 25: 15–18.<br />

Richard-Yegres N, Yegres, Zeppenfeldt G (1992). Cromomicosis: endemia rural,<br />

laboral y familiar en Venezuela. Revista Iberoamericana Micologia 9: 38–41.<br />

Riddley M (1957). <strong>The</strong> natural habitat of <strong>Cladosporium</strong> carrionii a cause of<br />

chromoblastomycosis. Australian Journal of Dermatology 4: 23–27.<br />

Rubin HA, Bruce S, Rosen T, McBride ME (1991). Evidence for percutaneous<br />

inoculation as the mode of transmission for chromoblastomycosis. Journal of<br />

the American Academy of Dermatology 25: 951–954.<br />

Salgado L, Pereira da Silva J, Picanço Diniz JÁ, Batista da Silva M, Fagundes da<br />

Costa P, Teixeira C, Salgado UI (2004). Isolation of Fonsecaea pedrosoi from<br />

thorns of Mimosa pudica, a probable natural source of chromoblastomycosis.<br />

Revista Instituto de Medicina Tropical, Sao Paulo 46: 33–36.<br />

Schoch CL, Shoemaker RA, Seifert KA, Hambleton A, Spatafora JW, Crous PW<br />

(2006). A multigene phylogeny of the Dothideomycetes using four nuclear loci.<br />

Mycologia 98: 1041–1052.<br />

Seifert KA, Nickerson NL, Corlett M, Jackson ED, Louis-Seize G, Davies RJ<br />

(2004). Devriesia, a new hyphomycete <strong>genus</strong> to accommodate heat-resistant,<br />

cladosporium-like fungi. Canadian Journal of Botany 82: 914–926.<br />

Silva JP, Rocha RM da, Moreno JS (1995). <strong>The</strong> coconut babaçu (Orbignya phalerata)<br />

as a probable risk of human infection by the agent of chromoblastomycosis<br />

in the State of Maranhão, Brazil. Revista do Sociedad Brasiliera de Medicina<br />

Tropical, Sao Paulo 28: 49–52.<br />

Swofford DL (1981). Utility of the distance-Wagner procedure. In: Advances in<br />

cladistics, vol. 1 (Funk VA <strong>and</strong> Brooks DR, eds.). Annals of the New York<br />

Botanical Gardens, New York: 25–44.<br />

Swofford DL (2003). PAUP*. Phylogenetic Analysis Using Parsimony (*<strong>and</strong> Other<br />

Methods). Version 4. Sinauer Associates, Sunderl<strong>and</strong>, Massachusetts.<br />

Szaniszlo PJ (2002). Molecular genetic studies of the model <strong>dematiaceous</strong><br />

pathogen Wangiella dermatitidis. International Journal of Medical Microbiology<br />

292: 381–390.<br />

Tamura K, Nei M (1993). Estimation of the number of nucleotide substitutuions in the<br />

control region of mitochondrial DNA in humans <strong>and</strong> chimpanzees. Molecular<br />

Biolology <strong>and</strong> Evolution 10: 512–526.<br />

Trejos A (1953). Cromoblastomycosis experimental en Bufo marinus. Revista de<br />

Biologia Tropical 1: 39–53.<br />

Trejos A (1954). <strong>Cladosporium</strong> carrionii n. sp. <strong>and</strong> the problem of cladosporia<br />

isolated from chromoblastomycosis. Revista de Biologia Tropical 2: 75–112.<br />

Vélez F, Chávez J (1980). Los Cactus de Venezuela. Colecciones Inagro,<br />

Venezuela.<br />

White TJ, Bruns T, Lee S, Taylor J (1990). Amplification <strong>and</strong> direct sequencing of<br />

fungal ribosomal RNA genes for phylogenetics. In: PCR Protocols: a guide to<br />

methods <strong>and</strong> applications (Innis MA, Gelf<strong>and</strong> DH, Sninsky JJ, White TJ, eds).<br />

Academic Press, San Diego, California: 315–322.<br />

Yegres F, Niel F, Gantier JC, Richard-Yegres N (1998). Murine humoral immune<br />

response against Cladophialophora carrionii <strong>and</strong> Fonsecaea pedrosoi infection.<br />

Journal de Mycologie Médical 8: 179–182.<br />

Yegres F, Richard-Yegres N (2002). Cladophialophora carrionii: Aportes al<br />

conocimiento de la endemia en Venezuela durante el siglo XX. Revista de la<br />

Sociedad Venezuelana de Microbiologίa 2: 153–157.<br />

Yegres F, Richard-Yegres N, Medina-Ruiz E, González-Vivas R (1985).<br />

Cromomicosis por <strong>Cladosporium</strong> carrionii en criadores de caprinos del estado<br />

Falcon. Investigaciόn Clínica 26: 235–246.<br />

Yegres F, Richard-Yegres N, Nishimura K, Miyaji M (1991). Virulence <strong>and</strong><br />

pathogenicity of human <strong>and</strong> environmental isolates of <strong>Cladosporium</strong> carrionii in<br />

new born ddY mice. Mycopathologia 114: 71–76.<br />

Yegres F, Richard-Yegres N, Perez-Blanco M (1996). Cromomicosis. In: Temas<br />

de Micologia Médica. (Bastardo de Albornoz M, ed.). Caracas, Venezuela:<br />

87–102.<br />

Yegüez-Rodriguez J, Richard-Yegres N, Yegres F, Rodríguez Larralde A (1992).<br />

Cromomicosis: susceptibilidad genética en grupos familiares de la zona<br />

endémica en Venezuela. Acta Cientίfica Venezuelana 43: 98–102.<br />

Zeng J-S, Sutton DA, Fothergill AW, Rinaldi MG, Harrak MJ, Hoog GS de (2007).<br />

Spectrum of clinically relevant Exophiala species in the U.S.A. Journal of<br />

Clinical Microbiology: in press.<br />

234


available online at www.studiesinmycology.org<br />

doi:10.3114/sim.2007.58.09<br />

Studies in Mycology 58: 235–245. 2007.<br />

Taxonomy, nomenclature <strong>and</strong> phylogeny of three cladosporium-like hyphomycetes,<br />

Sorocybe resinae, Seifertia azaleae <strong>and</strong> the Hormoconis anamorph of Amorphotheca<br />

resinae<br />

K.A. Seifert, S.J. Hughes, H. Boulay <strong>and</strong> G. Louis-Seize<br />

Biodiversity (Mycology & Botany), Eastern Cereal <strong>and</strong> Oilseed Research Centre, Agriculture & Agri-Food Canada, Ottawa, Ontario K1A 0C6 Canada<br />

*Correspondence: Keith A. Seifert, seifertk@agr.gc.ca<br />

Abstract: Using morphological characters, cultural characters, large subunit <strong>and</strong> internal transcribed spacer rDNA (ITS) sequences, <strong>and</strong> provisions of the International Code<br />

of Botanical Nomenclature, this paper attempts to resolve the taxonomic <strong>and</strong> nomenclatural confusion surrounding three species of cladosporium-like hyphomycetes. <strong>The</strong> type<br />

specimen of Hormodendrum resinae, the basis for the use of the epithet resinae for the creosote fungus {either as Hormoconis resinae or <strong>Cladosporium</strong> resinae) represents<br />

the mononematous synanamorph of the synnematous, resinicolous fungus Sorocybe resinae. <strong>The</strong> phylogenetic relationships of the creosote fungus, which is the anamorph of<br />

Amorphotheca resinae, are with the family Myxotrichaceae, whereas S. resinae is related to Capronia (Chaetothyriales, Herpotrichiellaceae). Our data support the segregation<br />

of Pycnostysanus azaleae, the cause of bud blast of rhododendrons, in the recently described anamorph <strong>genus</strong> Seifertia, distinct from Sorocybe; this species is related to the<br />

Dothideomycetes but its exact phylogenetic placement is uncertain. To formally stabilize the name of the anamorph of the creosote fungus, conservation of Hormodendrum<br />

resinae with a new holotype should be considered. <strong>The</strong> paraphyly of the family Myxotrichaceae with the Amorphothecaceae suggested by ITS sequences should be confirmed<br />

with additional genes.<br />

Key words: Amorphothecaceae, <strong>Cladosporium</strong> resinae, creosote fungus, Hormoconis resinae, jet fuel fungus, kerosene fungus, Myxotrichaceae, Pycnostysanus, resinicolous<br />

fungi.<br />

INTRODUCTION<br />

<strong>The</strong> ascomycete Amorphotheca resinae Parbery (1969) grows in<br />

hydrocarbon-rich substrates such as jet fuel, cosmetics <strong>and</strong> wood<br />

preserved with creosote or coal tar. This fungus is widely known<br />

by the anamorph name Hormoconis resinae (Lindau) Arx & G.A.<br />

de Vries or its obligate synonym <strong>Cladosporium</strong> resinae (Lindau)<br />

G.A. de Vries. It produces lightly pigmented, warty conidiophores,<br />

<strong>and</strong> branched, acropetally developing chains of lightly pigmented<br />

ameroconidia lacking conspicuous scars (Fig. 1B–E). This species<br />

is known colloquially as the “creosote fungus”, the “kerosene<br />

fungus” or the “jet fuel fungus”; to avoid confusion caused by the<br />

many heterotypic names with the epithet “resinae”, in this paper<br />

we generally will use the oldest of these informal names, “creosote<br />

fungus”, when referring to A. resinae or its anamorph. This fungus<br />

grows in jet fuel contaminated with small amounts of water, <strong>and</strong> the<br />

mycelium clogs fuel lines <strong>and</strong> corrodes metal parts. Consequently,<br />

fuel tanks in airports are monitored for this fungus by private<br />

companies using various physiological or biochemical tests.<br />

Sorocybe resinae (Fr.) Fr. produces dark black colonies on conifer<br />

resin, comprising dark synnemata <strong>and</strong> an effuse mononematous<br />

synanamorph, both with cladosporium-like conidiogenous cells<br />

<strong>and</strong> conidia. Unlike the anamorph of the creosote fungus, the<br />

conidia of Sorocybe resinae are dark brown <strong>and</strong> the lateral walls<br />

are conspicuously thicker than the poles (Fig. 2D–G). Colonies<br />

with only the mononematous anamorph sometimes occur, <strong>and</strong><br />

the mononematous anamorph can be sparse on colonies bearing<br />

synnemata. However, the conidia of the mononematous anamorph<br />

have identical pigmentation <strong>and</strong> lateral wall thickening to that of<br />

the synnematous anamorph. <strong>The</strong> mononematous anamorph rarely<br />

has been referred to by its own binomial name although, as we will<br />

show, there is a species epithet available. For the same reasons<br />

given above for Amorphotheca Parbery, generally we will refer to<br />

Sorocybe resinae herein as “the resin fungus”.<br />

Despite the micromorphological differences noted above, there<br />

is disagreement about whether the creosote fungus is conspecific<br />

with the mononematous synanamorph of the resin fungus (Parberry<br />

1969). <strong>The</strong> name for the anamorph of the creosote fungus is based<br />

on Hormodendrum resinae Lindau (1906). Christensen et al.<br />

(1942) presented a study of a cladosporium-like fungus commonly<br />

isolated from wood impregnated with creosote <strong>and</strong> coal tar <strong>and</strong><br />

applied Lindau’s name without examining its type. A later ecological<br />

study by Marsden (1954) employed the same name for the same<br />

fungus. An extra dimension was added to the confusion when de<br />

Vries (1952, using the name <strong>Cladosporium</strong> avellaneum G.A. de<br />

Vries) described four formae for the creosote fungus (differing in<br />

the colours of their conidia, the production of setae, or the total<br />

absence of conidia), each based on single conidium isolates made<br />

from one parent culture. De Vries (1955) <strong>and</strong> Parberry (1969)<br />

examined the holotype of Hormodendrum resinae <strong>and</strong> concluded<br />

that it represented the creosote fungus. Hughes (1958), prior to<br />

the description of Amorphotheca or Hormoconis Arx & G.A. de<br />

Vries, examined the same specimen <strong>and</strong> considered it to be the<br />

mononematous synanamorph of the resin fungus. If Hughes (1958)<br />

is correct, then neither the species Hormodendrum resinae, nor<br />

the <strong>genus</strong> that it typifies, Hormoconis, can represent the creosote<br />

fungus, as intended by Parberry (1969) or von Arx <strong>and</strong> de Vries (in<br />

von Arx 1973).<br />

In this paper, we present micromorphological, cultural <strong>and</strong><br />

molecular evidence that the resin fungus is a different species from<br />

the creosote fungus. Combined with re-examination of the holotype<br />

of Hormodendrum resinae, this information is used to provide a<br />

revised taxonomy <strong>and</strong> nomenclature for these two species. A third<br />

cladosporium-like fungus, Seifertia azaleae, is also considered in<br />

our discussion of generic concepts.<br />

Historical review<br />

<strong>The</strong> history of the fungus now known as Sorocybe resinae began<br />

with Fries (1815), who described Racodium resinae Fr. as follows:<br />

“310. Racodium resinae, expansum molliusculum dense contextum nigrum, filis<br />

inaequalibus.<br />

In resina Pini Abietis in silvis Suecia passim.<br />

Habitu et loco natali distinctum. Fila divaricato-ramosa; alia rigidula apice<br />

capituli sera, sub miscrosc. Coremio Link similia, Demat. villosum Schleich. huic<br />

simile; sed sub microsc. fila maxime differunt.”<br />

235


Seifert et al.<br />

Fig. 1. Amorphotheca resinae, colony characters <strong>and</strong> anamorph micromorphology. A. 10-d-old colony on PDA. B, D–E. Micromorphology of conidiophores, showing acropetal<br />

conidial chains, ramoconidia, <strong>and</strong> conidia. C. Conidia. DAOM 170427; for C, E see scale bar in D.<br />

<strong>The</strong> comparison with Coremium Link indicates the probability of a<br />

synnematous fungus, <strong>and</strong> an authentic specimen of Fries’ fungus,<br />

which as the only known authentic material we interpret as the<br />

holotype, is preserved in Link’s herbarium (see below). It represents<br />

the synnematous form of the resin fungus 1 .<br />

1 Persoon (1822) described a form of R. resinae “β piceum”. Hughes (1968) examined<br />

the holotype of this form, <strong>and</strong> it represents the mycelium of the ascomycete<br />

Strigopodia resinae (Sacc. & Bres.) S.J. Hughes. This taxon is thus not relevant to<br />

the three species that are the focus of this paper.<br />

Fries (1832) later transferred his species to Sporocybe<br />

Fr. (1825), a <strong>genus</strong> then used for relatively conspicuous dark<br />

hyphomycetes with dry spores (Mason & Ellis 1953). <strong>The</strong> 1832<br />

description explicitly stated... “capitulo rotundato inaequali, sporidiis<br />

seriatis, stipite aequali simplici.” <strong>The</strong> use of “capitulo” <strong>and</strong> “stipite”<br />

imply what would now be recognised as a synnematous fungus.<br />

Fries (1832) further characterised the habit of the fungus as “habitu<br />

stipitum Calicii,” a further comparison to a group of black, stipitate<br />

lichenized fungi classified in Calicium Pers., which under a h<strong>and</strong><br />

lens look <strong>similar</strong> to a dark synnematous fungus.<br />

Fries (1849) next described the <strong>genus</strong> Sorocybe Fr. for this<br />

fungus, as follows:<br />

236


Hormoconis resinae <strong>and</strong> morphologically <strong>similar</strong> taxa<br />

Fig. 2. Sorocybe resinae, synnematous form. A. Colony on bark of living, st<strong>and</strong>ing conifer. B. Synnemata. C. Four-month-old colony on DG18. D–G. Acropetally developing<br />

chains of conidia. Note that the lateral walls are conspicuously thickened; compare with Fig. 3. A, C. DAOM 239134. B, D–G. DAOM 11381.<br />

Sorocybe Fr.<br />

Habitus prioris. sed mycelium floccosum densum, stroma corneo-carbonaceum,<br />

sporis moniliformi-concatenatis basi excipulum incompletum praebens.<br />

1. S. resinae. Fr. 1–4. at raro fructif. Klotzsch. exs. C. 2.<br />

Because this description explicitly referred to the Systema, Fries<br />

presumably was segregating the fungus, originally described<br />

as Racodium resinae, into a new monotypic <strong>genus</strong> (McNeill et<br />

al. 2007; Art. 33.3) <strong>and</strong> this interpretation of R. resinae as the<br />

basionym generally has been followed in subsequent treatments<br />

of Sorocybe resinae.<br />

As noted in Table 1, Fries’ Racodium resinae was placed in<br />

several other hyphomycete genera by eighteenth century authors.<br />

<strong>The</strong>se diversions need not be reviewed in detail here because the<br />

modern status of these other genera, <strong>and</strong> their lack of <strong>similar</strong>ity with<br />

Sorocybe, is clear.<br />

Bonorden (1851) described Hormodendrum Bonord., with four<br />

species originally placed in Penicillium Link by Corda (1839); H.<br />

olivaceum (Corda) Bonord. (≡ Penicillium olivaceum Corda 1839)<br />

was designated as lectotype by Clements & Shear (1931). This<br />

<strong>genus</strong> was frequently, but incorrectly, spelled “Hormodendron”.<br />

Bonorden’s descriptions <strong>and</strong> illustrations are of variable quality by<br />

modern st<strong>and</strong>ards, <strong>and</strong> his herbarium is unknown (Stafleu et al.<br />

1995). Consequently the actual identities of the species Bonorden<br />

placed in Hormodendrum are unknown <strong>and</strong> Corda’s <strong>Cladosporium</strong><br />

olivaceum (Corda) Bonord. was dismissed in Penicillium monographs<br />

because the drawing shows branched conidial chains<br />

(Thom 1930), although the specimen has apparently not been<br />

re-examined. <strong>The</strong> generic name was used as a segregate for<br />

<strong>Cladosporium</strong> Link by some authors (e.g. Kendrick 1961), in<br />

particular for species with ameroconidia (de Vries 1952). Although it<br />

sometimes has been considered a synonym of <strong>Cladosporium</strong>, it will<br />

remain a nomen dubium until the type species is properly typified.<br />

Unaware of the resinicolous fungus described by Fries, Lindau<br />

described two species growing on conifer resin, Pycnostysanus<br />

resinae Lindau (1904), the type of this anamorph generic name,<br />

<strong>and</strong> Hormodendrum resinae Lindau (1906). <strong>The</strong> former was<br />

clearly illustrated <strong>and</strong> described as a synnematous species. <strong>The</strong><br />

protologue of the latter concludes with, “Mit Pycnostysanus resinae<br />

hat die Art nichts zu tun.” Clearly, Lindau observed no synnemata<br />

on the specimen of the mononematous fungus <strong>and</strong> he believed<br />

it was a different fungus, rather than what would now be called a<br />

synanamorph of the synnematous fungus that he had described<br />

previously. Lindau (1910) reproduced the 1904 illustration of<br />

Pycnostysanus resinae as Stysanus resinae (Fr.) Sacc. (1906),<br />

thus accepting its identity with the species originally described<br />

as Racodium resinae Fr. Lindau (1910) made no mention of<br />

Hormodendrum resinae, indicating he still made no association<br />

between the synnematous <strong>and</strong> mononematous fungi on resin.<br />

De Vries (1952) described a new species, <strong>Cladosporium</strong><br />

avellaneum G.A. de Vries, isolated from cosmetics. Later, he noted<br />

the <strong>similar</strong>ities between his C. avellaneum <strong>and</strong> the creosote fungus,<br />

<strong>and</strong> suggested that they were the same species (de Vries 1955),<br />

replacing the name of one of his previously described formae,<br />

i.e. viride, with the forma name resinae. He examined Lindau’s<br />

type of Hormodendrum resinae <strong>and</strong> decided that it provided an<br />

earlier epithet for C. avellaneum. He transferred the species into<br />

<strong>Cladosporium</strong> as C. resinae (Lindau) G.A. de Vries, <strong>and</strong> this name<br />

was widely used for the creosote fungus until 1973. This binomial<br />

is still commonly employed in non-taxonomic literature, especially<br />

commercial publications dealing with the creosote fungus.<br />

www.studiesinmycology.org<br />

237


Seifert et al.<br />

Table 1. Nomenclature <strong>and</strong> synonymies for the creosote fungus <strong>and</strong> the resin fungus, showing the use of the same basionym for the two fungi. <strong>The</strong> “false” names <strong>and</strong><br />

synonymies for the anamorph of the resin fungus are indicated by blue text. <strong>The</strong> second nomenclatural solution described in the text would have the effect of switching the<br />

blue text to black for the creosote fungus, <strong>and</strong> to simultaneously switch the equivalent black text to blue for the mononematous synanamorph of the resin fungus. Holotypes<br />

we have examined, <strong>and</strong> the herbarium where they are deposited, are marked with exclamation points, <strong>and</strong> details of these specimens are noted in Materials <strong>and</strong> Methods.<br />

Creosote fungus<br />

Teleomorph: Amorphotheca resinae Parberry, Australian J. Bot. 17: 340. 1969.<br />

Anamorph<br />

Hormodendrum resinae Lindau, in Rabenh. Krypt.-Fl., 2, 1 (Pilze) 8: 699. 1906 (B!).<br />

≡ <strong>Cladosporium</strong> resinae (Lindau) G.A. de Vries, Antonie van Leeuwenhoek 21: 167. 1955.<br />

≡ Hormoconis resinae (Lindau) von Arx & G.A. de Vries, in von Arx, Verh. K. Ned. Akad. Wet., Afd. Natuurk. 61: 62. 1973.<br />

= <strong>Cladosporium</strong> avellaneum G.A. de Vries, Contribution to the knowledge of the <strong>genus</strong> <strong>Cladosporium</strong>, Uitg. Druk. Holl<strong>and</strong>ia, p. 56, 1952.<br />

Resin fungus<br />

Mononematous synanamorph:<br />

Hormodendrum resinae Lindau, in Rabenh. Krypt.-Fl., 2, 1 (Pilze) 8: 699. 1906 (B!)<br />

≡ <strong>Cladosporium</strong> resinae (Lindau) G.A. de Vries, Antonie van Leeuwenhoek 21: 167. 1955.<br />

≡ Hormoconis resinae (Lindau) von Arx & G.A. de Vries, in von Arx, Verh. K. Ned. Akad. Wet., Afd. Natuurk. 61: 62. 1973.<br />

Synnematous anamorph:<br />

Sorocybe resinae (Fr.) Fr., Summa Veg. Scan. 2: 468. 1849.<br />

≡ Racodium resinae Fr., Obs. Mycol. 1: 216. 1815 (basionym) (B!).<br />

≡ Sporocybe resinae (Fr.) Fr., Syst. Mycol. 3: 341. 1832.<br />

≡ Dendryphion resinae (Fr.) Corda, Icon. Fung. 6: 11. 1854.<br />

≡ Stysanopsis resinae (Fr.) Ferr., Flora Ital. Crypt., 1 (Fungi, Hyphales), p. 187. 1910.<br />

? = Dematium nigrum Link, Mag. ges. naturf. Fr. 3: 21. 1809 (B!).<br />

≡ Sporotrichum nigrum (Link) Link, Mag. Ges. naturf. Fr. Berlin 7: 35. 1815.<br />

= Pycnostysanus resinae Lindau, Verh. Bot. Ver. Br<strong>and</strong>enb. 45 : 160. 1904 (B!).<br />

≡ Stysanus resinae (Lindau) Sacc., Syll. Fung. 18: 651. 1906.<br />

In his study of type collections of classical hyphomycetes,<br />

Hughes (1958) included Pycnostysanus resinae Lindau <strong>and</strong><br />

Hormodendrum resinae Lindau as facultative synonyms of<br />

Sorocybe resinae (Fr.) Fr., with Racodium resinae Fr. <strong>and</strong> several<br />

other nomenclatural variants as obligate synonyms (Table 1). <strong>The</strong><br />

synnematous Pycnostysanus resinae was cited as “Pycnostysanus<br />

state [i.e. synanamorph] of Sorocybe resinae”. Hormodendrum<br />

resinae thus remained to represent the mononematous<br />

synanamorph of what was interpreted as a single species.<br />

Parberry (1969) described a cleistothecial ascomycete,<br />

Amorphotheca resinae, for the teleomorph of the creosote fungus.<br />

He also examined the holotype of Hormodendrum resinae <strong>and</strong><br />

agreed with the conclusions of de Vries (1955). He used the<br />

epithet resinae for the teleomorph to correspond with that of the<br />

anamorph. He discounted the possibility that the synnematous<br />

Sorocybe resinae could be the same fungus as Hormodendrum<br />

resinae because synnemata never developed in his cultures of the<br />

creosote fungus.<br />

Von Arx <strong>and</strong> de Vries (in von Arx 1973) described the <strong>genus</strong><br />

Hormoconis, typified by Hormodendrum resinae, with the new<br />

combination Hormoconis resinae (Lindau) Arx & G.A. de Vries.<br />

<strong>The</strong>ir intention was to erect an anamorph <strong>genus</strong> for the anamorph<br />

of the creosote fungus, which they suggested was improperly<br />

classified in <strong>Cladosporium</strong> because it lacked darkened, thickened<br />

secession scars on the conidia.<br />

A third cladosporium-like fungus is relevant to this story.<br />

Seifertia azaleae (Peck) Partridge & Morgan-Jones [until recently<br />

known as Pycnostysanus azaleae (Peck) E.W. Mason] is a<br />

cosmopolitan fungus causing bud blast <strong>and</strong> twig blight of azaleas<br />

<strong>and</strong> rhododendrons. This species is morphologically <strong>similar</strong> to<br />

Sorocybe resinae, but the conidia are paler <strong>and</strong> lack laterally<br />

thickened walls. Sorocybe <strong>and</strong> Pycnostysanus have often been<br />

considered taxonomic synonyms (Ellis 1976, Carmichael et al.<br />

1980); as shown above, both are based on the synnematous form<br />

of the resin fungus. Partridge <strong>and</strong> Morgan-Jones (2002) argued that<br />

Sorocybe resinae <strong>and</strong> “Pycnostysanus azaleae” are not congeneric,<br />

<strong>and</strong> described the new <strong>genus</strong> Seifertia Part. & Morgan-Jones for<br />

the Rhododendron fungus. <strong>The</strong>y observed that the connection<br />

between conidia in Seifertia azaleae is much narrower than in<br />

Sorocybe resinae, <strong>and</strong> that minute denticles are visible on the<br />

conidiogenous cells of the former fungus. <strong>The</strong> broader connections<br />

between conidia of Sorocybe resinae result in broadly protuberant<br />

conidiogenous loci on the conidiogenous cells, <strong>and</strong> more truncate<br />

detached conidia.<br />

MATERIALS AND METHODS<br />

Herbarium material <strong>and</strong> fungal strains<br />

Full details of herbarium material examined are listed below.<br />

Cultures <strong>and</strong> dried herbarium specimens were studied in 90 %<br />

lactic acid without stains; preparations of some exsiccate <strong>and</strong> types<br />

were mounted in glycerin jelly. Cultures were grown on potatodextrose<br />

agar (PDA, Difco), oatmeal agar (OA, Samson et al.<br />

2004), Blakeslee’s malt extract agar (MEA, Samson et al. 2004) <strong>and</strong><br />

dichloran-18 % glycerol agar (DG-18, Samson et al. 2004). Colony<br />

characters were taken from cultures grown at 25 °C in darkness.<br />

Cultures are maintained in the Canadian Collection of Fungal<br />

Cultures (DAOM), Agriculture & Agri-Food Canada, Ottawa.<br />

238


Hormoconis resinae <strong>and</strong> morphologically <strong>similar</strong> taxa<br />

Exsiccati <strong>and</strong> types<br />

Dematium nigrum [scr. Link]. E. Hbr. Link (23) = Sporocybe<br />

resinae III. 341 [scr. ? ] (herb. Link, B).<br />

Hormodendrum resinae Lindau, n. sp. Fl. v. Hamburg 206, auf<br />

Harz an Picea excelsa, Sachsenwald, leg. O. Jaap, 29-4-1906.<br />

[scr. Lindau]. (DAOM 41888, slide prepared from the holotype<br />

preserved in B.)<br />

Pycnostysanus resinae Lindau nov. gen. et nov. spec., Kabát et<br />

Bubák: Fungi imperfecti exsiccati no. 99. Auf erhärteten Fichtenharz<br />

an Brockenweg, am Dreieckigen Pfahl in Harz, Deutschl<strong>and</strong>, leg.<br />

G. Lindau, 13.VIII. 1903 (holotype, B).<br />

Racodium resinae Fries. E. Hbr. Link, Fries legi, Smol. [scr.<br />

Fries]. (DAOM 41890, slide prepared from herb. Link, B). This<br />

is the presumed holotype of R. resinae, the basionym for the<br />

resin fungus, Sorocybe resinae. <strong>The</strong> specimen includes dark,<br />

decapitated synnemata, brown conidia with laterally thickened<br />

walls, <strong>and</strong> acropetal conidial chains, allowing it to be recognised<br />

as the fungus we now know as S. resinae. Fries perhaps sent this<br />

fungus to Link to see if it could be differentiated from Coremium<br />

Link. <strong>The</strong> minimal details, that the fungus was collected by Fries,<br />

presumably in Smål<strong>and</strong> (a province of Sweden), match the details<br />

in the protologue of this species.<br />

Sorocybe resinae. “Fungi Rhenani Fasc. II, 1863, L. Fuckel, no.<br />

129, ad Abietis resinam, raro Hieme, in sylva Hostrichiensi” (as<br />

Myxotrichum resinae Fr., DAOM 55543 ex FH). “Flora Suecica,<br />

2956, Ad resinam piceae, Smål<strong>and</strong>: Femsjö, Prostgaidsshogen,<br />

6 Aug. 1929, leg. J.A. Nannfeldt, s.n.” (as Stysanus resinae (Fr.)<br />

Sacc., DAOM 41891 ex UPS). “Flora Suecica, 4709, Ad resinam<br />

abietinum, Uppl<strong>and</strong>: Bondkysko sin Valsätra, 9 May 1932, leg. J.A.<br />

Nannfeldt” (as Hormodendrum resinae Lindau, DAOM 41889 ex<br />

UPS). “[ on wood scr. Berkeley] J.E. Vize, Hereford 1877” (as Torula<br />

pinophila Fr., DAOM 113425 ex K). “Sydow, Mycotheca germanica,<br />

350. Auf Fichtenharz… am Brockenweg 30.9.1904, leg. P. Sydow”<br />

(DAOM 41893).<br />

Other material examined<br />

Sorocybe resinae. Canada, British Columbia: Burnaby, Central Park, on resin of<br />

Tsuga heterophylla, leg. S. & L. Hughes, 17 Aug. 2000 (DAOM 228572a, 228573a);<br />

Cameron Lake, Cathedral Grove, on Pseudotsuga menziesii, leg. isol. S.J. Hughes,<br />

21 Aug. 1957 (DAOM 56088a). Ladysmith, Ivy Green Park, on resinous exudates,<br />

leg. R.J. B<strong>and</strong>oni no. BC-978, 18 Apr. 1960, det. S.J. Hughes (DAOM 70462).<br />

North Vancouver, Lynn Valley Conservation Area, leg. det. S.J. Hughes, 1 Jul. 1975<br />

(DAOM 139385); North Vancouver, Lynn Valley Conservation Area, on bark of living<br />

conifer (probably Pseudotsuga menziesii), leg. isol. K.A. Seifert no. 1574, 26 May<br />

2002 (single conidium isolate, culture <strong>and</strong> specimen DAOM 239134; ITS GenBank<br />

EU030275, LSU GenBank EU030277); Terrace, near Kalum, on Tsuga heterophylla,<br />

leg. W.G. Ziller no. V-6549, 10 July 1950, det. S.J. Hughes (DAOM 59657); Queen<br />

Charlotte Isl<strong>and</strong>s, east coast of Moresby Isl<strong>and</strong>, north side of Gray Bay, 53°08’ N,<br />

131°47’ W, on Picea sitchensis, leg. I. Brodo, M.J. Schepanek, W.B. Schofield,<br />

28 Sep. 1973, det. S.J. Hughes (DAOM 144757); Queen Charlotte Isl<strong>and</strong>s,<br />

Graham Isl<strong>and</strong>, Tow Hill area, on resin of Picea sitchensis, leg. S.A. Redhead no.<br />

4440, 20 Sep. 1982, det. G.P. White (DAOM 184025); Revelstoke, Wigwam, on<br />

Tsuga heterophylla, leg. W. Ziller V-6567 det. S.J. Hughes, 6 Jun. 1950 (DAOM<br />

59710); Vancouver Isl<strong>and</strong>, Cathedral Grove, Cameron Lake, on Pseudotsuga<br />

menziesii, leg. det. S.J. Hughes, 21 Aug. 1957 (DAOM 56088a); Vancouver Isl<strong>and</strong>,<br />

Caycuse, on resin of Pseudotsuga menziesii, leg. det. S.J. Hughes, 17 Jul. 1972<br />

(DAOM 139355); Vancouver Isl<strong>and</strong>, Lake Cowichan, Honeymoon Bay, on resin<br />

of Pseudotsuga menziesii, leg. J Ginns, det. S.J. Hughes, 29 Oct. 1971 (DAOM<br />

134968); Vancouver Isl<strong>and</strong>, Lake Cowichan, Mesachie Lake Forest Experimental<br />

Station, leg. det. S.J. Hughes, 5 Jul. 1972 (DAOM 139277a, DAOM 139278) <strong>and</strong><br />

6 Jul. 1072 (DAOM 139281). Czechoslovakia, Ještěd near Liberec, leg. det. S.J.<br />

Hughes, on resin of Larix europaea, 10 May 1955 (DAOM 51723). United States,<br />

Oregon: Andrews’ Experimental Forest, Forest Service Rd. no 1553, on resin of<br />

Tsuga heterophylla, leg. det. S.J. Hughes, 10 May 1969 (DAOM 134565); Andrews’<br />

Experimental Forest, Blue River, on resin of conifer, cut wood, leg. det. K.A. Seifert<br />

no. 69, 10 Jul. 1981 (DAOM 228203); Oregon, del Norte Co., J. Smith’s State Park,<br />

on Tsuga heterophylla, leg. det. S.J. Hughes, 11 May 1069 (DAOM 134614); Devil’s<br />

Elbow State Park, Cape Perpetus, on Picea sitchensis, leg. det. S.J. Hughes, 6 May<br />

www.studiesinmycology.org<br />

1969 (DAOM 134615); Linn Co., near Cascadia, on Pseudotsuga menziesii, leg. R.<br />

Fogel, det. S.J. Hughes, 14 May 1969 (DAOM 127885); U.S. Forest Service Rd. no.<br />

126, North fork Cape Creek, on resin of Abies gr<strong>and</strong>is, leg. det. S.J. Hughes, 7 May<br />

1969 (DAOM 134852,134563); Willamette National Forest, McKenzie Bridge Camp<br />

Grounds, leg. det. S.J. Hughes, 10 May 1969 (DAOM 134564). Washington: Kittitas<br />

Co., Wanatchee National Forest, Rocky Run, on Abies nobilis, leg. Field Mycology<br />

Class 1955, 22 Jul. 1955, det. S.J. Hughes, (mononematous synanamorph only,<br />

DAOM 118934 ex WSP 45210, as Helminthosporium sp.); Jefferson Co., Olympic<br />

National Forest, 10 mi Camp, Sec. 17, T26N, R3W, on Pseudotsuga mucronata,<br />

leg. Field Mycology Class, 22 Jul. 1955 (DAOM 113801 ex WSP 45212, as<br />

Helminthosporium); Grays Harbor Co., Twin Harbors Beach State Park, resin of<br />

Picea sitchensis, leg. W.B. & V.G. Cooke, 24 Jul. 1951, det. S.J. Hughes (DAOM<br />

118970 ex WSP 28432).<br />

Amorphotheca resinae. Isolated from jet fuel by P. Emonds (culture, DAOM<br />

170427 = ATCC 22711, ITS GenBank EU030278, LSU GenBank EU030280).<br />

Canada, British Columbia, source unknown, isol. “Mrs. Volkoff”, Jul. 1969 (culture,<br />

DAOM 194228, ITS GenBank EU030279).<br />

Seifertia azaleae. All on flower buds of Rhododendron spp. Canada, British<br />

Columbia: Burnaby, Central Park, leg. S.J. Hughes, 17 Aug. 2000 (DAOM 228571);<br />

Vancouver, Stanley Park, leg. K.A. Seifert no. 1571, 11 May 2002 (culture <strong>and</strong><br />

specimen, DAOM 239136, LSU GenBank EU030276). Irel<strong>and</strong>, Munter, Kerry,<br />

near Glenbeigh (ca. N 52° 03’ W 9° 54’), leg. K.A. Seifert no. 3197, 26 Sep. 2006<br />

(culture <strong>and</strong> specimen, DAOM 239135, ITS GenBank EU030273). Netherl<strong>and</strong>s,<br />

Gelderl<strong>and</strong>, Kröller-Müller Museum, leg. K.A. Seifert no. 1235, 12 May 2000 (DAOM<br />

227136). United Kingdom, Wales, Hafod Estate (ca. N 52° 22’ W 3° 51’), leg. K.A.<br />

Seifert no. 3198, 1 Oct. 2006 (culture <strong>and</strong> specimen, DAOM 239137, ITS GenBank<br />

EU030274).<br />

DNA extraction, amplification <strong>and</strong> sequencing<br />

DNA was isolated using a FastDNA Kit <strong>and</strong> the FastPrep<br />

FP120 (BIO 101 Inc.) or an UltraClean Microbial DNA Isolation<br />

Kit (Mo Bio Laboratories, Inc., Solana Beach, CA, U.S.A.) using<br />

mycelium removed from agar cultures. PCR <strong>and</strong> cycle sequencing<br />

reactions were performed on a Techne Genius thermocycler<br />

(Techne Cambridge Ltd.). PCR reactions were performed using<br />

Ready-To-Go Beads (Amersham Canada Ltd.) in 25 µL volumes,<br />

each containing 20–100 ng of genomic DNA, 2.5 units pure Taq<br />

DNA Polymerase, 10 mM Tris-HCl (pH 9.0), 50 mM KCl, 1.5 mM<br />

MgCl 2<br />

, 200 µM of each dNTP, 0.2 µL of each primer (50 µM), <strong>and</strong><br />

stabilizers including bovine serum albumin. <strong>The</strong> reaction profile<br />

included an initial denaturation for 4 min at 94 °C, followed by 30<br />

cycles of 1.5 min denaturation at 95 °C, 1 min annealing at 56 °C,<br />

2 min extension at 72 °C, with a final extension of 10 min at 72 °C.<br />

Amplicons were purified by ethanol/sodium acetate precipitation<br />

<strong>and</strong> resuspended as recommended for processing on an ABI<br />

PRISM 3100 DNA Analyzer or an ABI 373 Stretch DNA Sequencer<br />

(Applied Biosystems, Foster, CA). Amplification products were<br />

sequenced using the BigDye v. 2.0 Terminator Cycle Sequencing<br />

Ready Reaction Kit (ABI Prism/Applied Biosystems) following the<br />

manufacturer’s directions. An approximately 1 000 bp portion of the<br />

large subunit (LSU) ribosomal DNA was amplified <strong>and</strong> sequenced<br />

using primers LR0R <strong>and</strong> LR6, <strong>and</strong> cycle-sequenced using primers<br />

LR0R, LR3R, LR16 <strong>and</strong> LR6 (Vilgalys & Hester 1990, Rehner &<br />

Samuels 1995; www.biology.duke.edu/fungi/mycolab/primers.<br />

htm). <strong>The</strong> complete ITS <strong>and</strong> 5.8S rRNA genes were amplified<br />

<strong>and</strong> sequenced using the primers ITS5 <strong>and</strong> ITS4, with ITS2 <strong>and</strong><br />

ITS3 primers used for cycle sequencing when necessary (White<br />

et al. 1990). Some sequences were derived from single PCR<br />

amplifications of the ITS5–LR6 region.<br />

Data matrices were subjected to parsimony analysis using<br />

heuristic searches in PAUP* v. 4.0b10 (Swofford 2002) with<br />

simple stepwise addition of taxa, <strong>and</strong> tree bisection-reconnection<br />

(TBR) branch swapping. Uninformative characters were removed<br />

for all analyses. Strict consensus trees were calculated, <strong>and</strong> the<br />

robustness of the phylogenies was tested using full bootstrap<br />

analyses (1 000 replications). For all analyses, GenBank accession<br />

numbers are given on the tree figures, <strong>and</strong> the sequences generated<br />

in this study are indicated in bold.<br />

239


Seifert et al.<br />

Fig. 3. Hormodendrum resinae, A–B. Conidiophores <strong>and</strong> acropetally developing chains of conidia. C. Conidia. Note that the lateral walls are conspicuously thickened compared<br />

to the walls at the poles. From a slide (DAOM 41888) prepared from the holotype (B).<br />

Byssoascus striatisporus U17912<br />

99<br />

96<br />

10 changes<br />

88 88<br />

99 “Mycosphaerella mycopappi” U43480<br />

* Seifertia azaleae DAOM 239136 EU030276<br />

Pleomassaria siparia AY004341<br />

*<br />

Pleospora herbarum AF382386<br />

* Cochliobolus heliconiae AF163978<br />

Setosphaeria monoceras AY016368<br />

Trematosphaeria heterospora AY016369<br />

Iodosphaeria aquatica AF452044<br />

Westerdykella cylindrica AY004343<br />

Letendraea helminthicola AY016362<br />

Lojkania enalia AY016363<br />

Botryosphaeria ribis AY004336<br />

Capnodium citri AY004337<br />

Discosphaerina fagi AY016359<br />

Ceramothyrium carniolicum AY004339<br />

Fonsecaea pedrosoi AF356666<br />

80 Phialophora americana AF050279<br />

99 79 Capronia mansonii AY004338<br />

Glyphium elatum AF346420<br />

97 Terfezia gigantea AF499449<br />

Sorocybe resinae DAOM 239134 EU030277<br />

Ascosphaera apis AY004344<br />

Saccharomyces cerevisiae J01355<br />

Phaeotrichum benjaminii AY004340<br />

Calicium viride AF356670<br />

Pertusaria mammosa AY212831<br />

73 Pseudocyphellaria perpetua AF401954<br />

Xylographa vitiligo AY212849<br />

Neofabraea alba AY064705<br />

Melanelia exasperatula AJ421436<br />

* Lophodermium pinastri AY004334<br />

99 Cudonia lutea AF433138<br />

Spathularia flavida AF433141<br />

Ascocoryne cylichnium AF353580<br />

Bisporella citrina AF335454<br />

Hormoconis resinae DAOM 170427 EU030280<br />

Rhytisma acerinum AF356696<br />

Pleuroascus nicholsonii AF096196<br />

Golovinomyces cichoracearum AB022360<br />

Fig. 4. Parsimony analysis of large subunit sequences, demonstrating the phylogenetic positions of Amorphotheca resinae, Sorocybe resinae <strong>and</strong> Seifertia azaleae (all shown<br />

in bold) in the Ascomycota. One of 12 equally parsimonious trees (1 888 steps, CI = 0.390, RI = 0.554, RC = 0.216, HI = 0.610) with Golovinomyces cichoracearum as the outgroup.<br />

Bootstrap values above 70 % are shown at the relevant nodes, with an asterisk representing 100 % bootstrap support; branches with thick lines occurred in all equally<br />

parsimonious trees.<br />

Leotiomycetes Dothideomycetes<br />

240


Hormoconis resinae <strong>and</strong> morphologically <strong>similar</strong> taxa<br />

<strong>The</strong> large subunit matrix was assembled from the closest<br />

BLAST matches using our sequences for the three fungi of<br />

interest, S. resinae, A. resinae <strong>and</strong> S. azaleae; Golovinomyces<br />

cichoracearum was added as an out-group to root the tree.<br />

Although these sequences were put into a single matrix, there<br />

is no implication that this data set represents the diversity of the<br />

Ascomycota. <strong>The</strong> alignment was calculated using MAFFT (Katoh et<br />

al. 2002) <strong>and</strong> adjusted using Se-Al (Sequence Alignment Program<br />

v. 1.d1; http://evolve.zoo.ox.ac.uk/software/Se-Al/main.html) to<br />

maximise homology.<br />

<strong>The</strong> internal transcribed spacers alignment including Sorocybe<br />

resinae was derived from an alignment of Capronia <strong>and</strong> related<br />

anamorphs used by Davey & Currah (2007), originally produced<br />

using MAFFT. This data set was modified considerably using Se-<br />

Al to maximise homology, but still included several areas where<br />

the homology of aligned sequences was difficult to evaluate. ITS<br />

sequences of Amorphotheca resinae were used to retrieve closely<br />

related sequences using a BLAST search of GenBank, <strong>and</strong> these<br />

relevant sequences were added to an alignment of Oidiodendron<br />

Robak sequences from the study of Hambleton et al. (1998), <strong>and</strong><br />

then adjusted using Se-Al.<br />

We attempted direct PCR from two specimens containing only<br />

the putative mononematous synanamorph of Sorocybe resinae<br />

(DAOM 228772a, 228573a), to allow comparison of sequences<br />

obtained from cultures of the synnematous synanamorph.<br />

<strong>The</strong>se attempts, using the same methods outlined above, were<br />

unsuccessful.<br />

RESULTS<br />

Cultural characters <strong>and</strong> micromorphology<br />

Most micromorphological characters of the resin fungus Sorocybe<br />

resinae (Partridge & Morgan-Jones 2002), the creosote fungus<br />

Amorphotheca resinae (Parbery 1969, de Vries 1952, 1955, Ho<br />

et al. 1999) <strong>and</strong> the rhododendron fungus Seifertia azaleae (Ellis<br />

1976, Partridge & Morgan-Jones 2002, Glawe & Hummel 2006) are<br />

well-described in the literature <strong>and</strong> will not be repeated here.<br />

<strong>The</strong> three species are readily distinguished based on growth<br />

rates <strong>and</strong> overall cultural phenotypes. Agar colonies of Sorocybe<br />

resinae are coal-black, wrinkled, <strong>and</strong> restricted in growth, no matter<br />

what agar medium is employed; even after 3 mo, the colonies are<br />

rarely more than 2 cm diam (Fig. 2C). Synnemata did not form in<br />

our cultures; in vivo, the synnemata produce branched, acropetal<br />

chains of conidia with laterally thickened walls (Figs 2D–G). No<br />

thickened, refractive or darkened secession scars were evident<br />

on individual conidia or ramoconidia. Conidial masses were<br />

removed from the mononematous <strong>and</strong> synnematous parts of a<br />

freshly collected specimen (DAOM 56088a) <strong>and</strong> grown on PDA<br />

<strong>and</strong> sterilised conifer wood. <strong>The</strong>re were no discernable differences<br />

between colonies derived from the two types of conidiophores, in<br />

all cases yielding restricted black colonies, or in their microscopic<br />

characters. <strong>The</strong>refore, we conclude that these two types of<br />

conidiophores represent synanamorphs of one fungus. An identical<br />

conclusion was reached by Partridge & Morgan-Jones (2002). We<br />

documented the occurrence of this fungus in California, Oregon,<br />

<strong>and</strong> Washington State, U.S.A. <strong>and</strong> British Columbia, Canada, on<br />

resinous exudates on Abies nobilis, Picea sitchensis, Pseudotsuga<br />

menziesii <strong>and</strong> Tsuga heterophylla.<br />

Microscopic features from the holotype specimen of<br />

Hormodendrum resinae Lindau are shown in Fig. 3. Dark, thickwalled<br />

conidiophore stipes give rise to branched, acropetally<br />

www.studiesinmycology.org<br />

developing conidial chains. <strong>The</strong> conidia are relatively darkly<br />

pigmented, <strong>and</strong> the lateral walls are more conspicuously thickened<br />

<strong>and</strong> darkened than the polar walls. <strong>The</strong>re are no obvious thickened,<br />

refractive or darkened secession scars on any of the cells. Apart<br />

from the production of synnemata, the characters of the conidia<br />

<strong>and</strong> conidium ontogeny are identical in Lindau’s specimen <strong>and</strong> the<br />

synnematous specimens of Sorocybe resinae examined.<br />

In contrast, both the resin fungus <strong>and</strong> the rhododendron fungus<br />

have spreading rather than restricted agar colonies. Cultures of<br />

the resin fungus are s<strong>and</strong>y brown (Kornerup & Wanscher 1989),<br />

planar <strong>and</strong> powdery, growing 4–4.5 cm diam in 10 d on PDA (Fig.<br />

1A). Cultures of the rhododendron fungus are slower, growing<br />

2.5–3.5 cm diam after 21 d on MEA (not shown). <strong>The</strong>y are planar<br />

<strong>and</strong> greyish brown, with an orange-brown reverse. No synnemata<br />

were observed in our cultures of the rhododendron fungus on MEA,<br />

OA or PDA, but cladosporium-like conidiation occurred in the aerial<br />

mycelium.<br />

Phylogeny<br />

<strong>The</strong> large subunit analysis (LSU) was used to demonstrate the<br />

general phylogenetic relationships of the resin fungus Sorocybe<br />

resinae (DAOM 239134), the creosote fungus Amorphotheca resinae<br />

(DAOM 170427, 194228) <strong>and</strong> the rhododendron fungus Seifertia<br />

azaleae (DAOM 239136), <strong>and</strong> subsequent analyses of the internal<br />

transcribed spacers were used to estimate more precise affinities.<br />

Fig. 4 shows the LSU analysis <strong>and</strong> demonstrates that Sorocybe<br />

resinae appears to be a member of the Herpotrichiellaceae,<br />

Chaetothyriales, A. resinae is related to the inoperculate<br />

discomycetes (Leotiomycetes) <strong>and</strong> Seifertia azaleae is most closely<br />

related to a sequence labelled Mycosphaerella mycopappi A. Funk<br />

& Dorworth, which is unrelated to Mycosphaerella s. str.<br />

For the ITS alignment of Sorocybe resinae, two preliminary<br />

parsimony analyses were conducted, one with informative<br />

characters from the full alignment, the second with a subset with<br />

179 characters excluded from seven ambiguously aligned regions.<br />

<strong>The</strong> consistency indices (full 0.301, partial 0.324), tree topologies,<br />

<strong>and</strong> bootstrap supports for the two analyses were relatively <strong>similar</strong>.<br />

<strong>The</strong>refore, the complete alignment was used for the tree presented<br />

here (Fig. 5). <strong>The</strong> data matrix included 57 taxa, with 352 of 752<br />

characters phylogenetically informative. Sorocybe resinae clearly<br />

is related to Capronia <strong>and</strong> allied anamorph genera, as suggested<br />

by the LSU analysis. In the ITS analysis (Fig. 6) it forms a wellsupported<br />

clade with C. villosa Samuels, that is a well-supported<br />

sister group to species now in three different anamorph genera,<br />

Phaeococcomyces nigricans (M.A. Rich & A.M. Stern) de Hoog,<br />

Ramichloridium cerophilum, <strong>and</strong> an undescribed species of<br />

Heteroconium Petr.<br />

<strong>The</strong> ITS matrix for A. resinae included 42 taxa, with 171<br />

phylogenetically informative characters in the 530 base alignment.<br />

<strong>The</strong> phylogenetic analysis confirmed the relationship of this<br />

species with the Leotiomycetes, <strong>and</strong> provided a more precise<br />

hypothesis of its family-level relationships (Fig. 6). Amorphotheca<br />

resinae DAOM 170427 <strong>and</strong> 194228 had identical ITS sequences<br />

to another strain of the same species reported in GenBank<br />

(AY251067, from Braun et al. 2003), <strong>and</strong> one bp substitution from<br />

a second strain (AF393726 based on the isotype ATCC 200942<br />

= <strong>CBS</strong> 406.68). <strong>The</strong>se four sequences formed a sister group<br />

to two sequences of “<strong>Cladosporium</strong>” breviramosum Morgan-<br />

Jones (AF393683, AF393684). <strong>The</strong> well-supported clade of A.<br />

resinae <strong>and</strong> C. breviramosum, which represent the proposed<br />

family Amorphothecaceae, was previously noted by Braun et<br />

al. (2003). <strong>The</strong> nesting of this clade within two well-supported<br />

241


Seifert et al.<br />

100<br />

76<br />

99<br />

99<br />

100<br />

AY064704 Neofabraea alba<br />

AF281377 Neofabraea alba<br />

AF281397 Neofabraea perennans<br />

AF281388 Neofabraea malicorticis<br />

AF281365 Neofabraea krawtzewii<br />

AY129289 Pseudeurotium ovale var. ovale<br />

AY129290 Pseudeurotium ovale var. milkoi<br />

AY129287 Pseudeurotium bakeri<br />

AY129286 Pseudeurotium zonatum<br />

AY129288 Pseudeurotium desertorum<br />

AF307760 Geomyces pannorum<br />

AJ390390 Geomyces asperulatus<br />

AY789290 Heyderia abietis<br />

Z81441 Piceomphale bulgarioides<br />

AY781230 Leptodontidium elatius<br />

AY348594 Calycina herbarum<br />

AF062802 Oidiodendron periconioides<br />

99<br />

100<br />

79<br />

99<br />

88<br />

100<br />

100<br />

89<br />

88<br />

100<br />

98<br />

100<br />

AF062787 Oidiodendron pilicola<br />

AF062798 Oidiodendron maius<br />

AF062790 Oidiodendron citrinum<br />

AF062789 Oidiodendron chlamydosporicum<br />

AF062797 Oidiodendron griseum<br />

AF307774 Oidiodendron tenuissimum<br />

AF062792 Oidiodendron flavum<br />

AF062810 Myxotrichum arcticum<br />

AF062809 Oidiodendron truncatum<br />

AF062808 Oidiodendron tenuissimum<br />

AF062805 Oidiodendron setiferum<br />

AF062788 Oidiodendron cerealis<br />

AJ635314 Oidiodendron myxotrichoides<br />

AF062811 Myxotrichum cancellatum<br />

AF062817 Byssoascus striatisporus<br />

AF062803 Oidiodendron rhodogenum<br />

AF062814 Myxotrichum deflexum<br />

88<br />

A. resinae DAOM 170427, 194228, EU030278-9<br />

AF393726 Amorphotheca resinae isotype<br />

AY251067 Amorphotheca resinae<br />

AF393684 <strong>Cladosporium</strong> breviramosum<br />

AF393683 <strong>Cladosporium</strong> breviramosum<br />

AF062813 Myxotrichum chartarum<br />

AF062812 Myxotrichum carminoparum<br />

AF062816 Myxotrichum stipitatum<br />

5 changes<br />

Fig. 5. Parsimony analysis of internal transcribed spacers sequences, demonstrating the position of Amorphotheca resinae (shown in bold) in the ascomycete family<br />

Myxotrichaceae. One of 44 equally parsimonious trees (645 steps, CI = 0.460, RI = 0.758, RC = 0.349, HI = 0.540) with mid-point rooting. Bootstrap values above 70 % are<br />

shown at the relevant nodes; branches with thick lines occurred in all equally parsimonious trees.<br />

Myxotrichaceae Pseudeurotiaceae<br />

clades of Myxotrichum spp. <strong>and</strong> the associated anamorph <strong>genus</strong><br />

Oidiodendron, which comprise the family Myxotrichaceae, has not<br />

been documented previously.<br />

<strong>The</strong> ITS sequences of two strains of Seifertia azaleae were<br />

474 bp <strong>and</strong> differed by one bp. BLAST searches with these<br />

sequences revealed significant homologies only with unidentified<br />

fungi, <strong>and</strong> lower probability matches with various members of<br />

the Dothideomycetes. <strong>The</strong>refore, no taxonomically meaningful<br />

phylogenetic analysis can be presented with these ITS sequences.<br />

<strong>The</strong> species does seem to have affinities with the Dothideomycetes,<br />

but the putative relationship with Mycosphaerella, suggested by the<br />

LSU analysis, could not be confirmed with the ITS analysis.<br />

DISCUSSION<br />

Micromorphological comparisons, differences in culture characters,<br />

<strong>and</strong> phylogenetic analysis all support the conclusion that the<br />

mononematous synanamorph of Sorocybe resinae, the resin<br />

fungus, is different from the anamorph of Amorphotheca resinae,<br />

the creosote fungus. Based on ribosomal DNA sequences, the<br />

creosote fungus is related to the family Myxotrichaceae, the <strong>genus</strong><br />

Myxotrichum <strong>and</strong> its Oidiodendron anamorphs (Fig. 5). In this<br />

gene tree, Myxotrichum <strong>and</strong> the Myxotrichaceae are paraphyletic,<br />

with Amorphotheca <strong>and</strong> the Amorphothecaceae nested within<br />

them. Sorocybe appears to be an additional anamorph <strong>genus</strong><br />

phylogenetically associated with Capronia (Herpotrichiellaceae,<br />

Chaetothyriales, Fig. 6). <strong>The</strong> genetic connection between the<br />

synnematous <strong>and</strong> mononematous morphs of S. resinae was<br />

verified by morphological comparison of polyspore isolates derived<br />

from the two synanamorphs. However, the living cultures are no<br />

longer available <strong>and</strong> the connection was not confirmed with single<br />

conidium isolations. <strong>The</strong> type specimen of Hormodendrum resinae<br />

(Fig. 3) is the basis for the application of the most frequently<br />

used anamorph epithet for the creosote fungus. This specimen<br />

represents the mononematous synanamorph of Sorocybe resinae,<br />

not the anamorph of Amorphotheca resinae.<br />

It is difficult to underst<strong>and</strong> how these two fungi were confused<br />

when their micromorphologies are so different. <strong>The</strong> conidia are of<br />

the same general size <strong>and</strong> shape, but in both morphs of Sorocybe<br />

resinae (Figs 2D–G, 3C), the lateral walls are conspicuously<br />

thickened, a condition not present in the creosote fungus (Fig. 1C),<br />

<strong>and</strong> the conidia are much darker. In his monograph of <strong>Cladosporium</strong>,<br />

de Vries (1952) noted that single conidium isolates of C. avellaneum<br />

gave rise to four different colony types. In 1955, he extended these<br />

observations <strong>and</strong> decided that the much darker resin fungus was<br />

242


Fig. 6<br />

Hormoconis resinae <strong>and</strong> morphologically <strong>similar</strong> taxa<br />

87<br />

89<br />

74<br />

100<br />

100<br />

83<br />

93<br />

94<br />

100<br />

93<br />

Capronia epimyces AY156968<br />

Capronia epimyces AF050245<br />

Capronia mansonii AF050247<br />

Capronia munkii AF050250<br />

Exophiala dermatitidis AY857526<br />

Exophiala heteromorpha AY857524<br />

Exophiala heteromorpha AF083207<br />

88<br />

97<br />

94<br />

Rhinocladiella similis AY857529<br />

Melanchlenus oligospermus AY163555<br />

Exophiala oligosperma AY857534<br />

Exophiala nishimurae AY163560<br />

Rhinocladiella basitona AY163561<br />

94<br />

Exophiala spinifera AM176734<br />

Ramichloridium mackenziei AY857540<br />

Rhinocladiella atrovirens AY618683<br />

Melanchlenus eumetabolus AY163554<br />

Capronia dactylotricha AF050243<br />

Ramichloridium anceps DQ826740<br />

Exophiala lecaniicorni AY857528<br />

Exophiala mesophila AF542377<br />

Exophiala attenuata AF549446<br />

Exophiala pisciphila DQ826739<br />

Exophiala crusticola AM048755<br />

Capronia nigerrima AF050251<br />

Capronia parasitica AF050253<br />

Capronia pulcherrima AF050256<br />

Capronia pilosella DQ826737<br />

Capronia pilosella AF050255<br />

Capronia coronata AF050242<br />

Cladophialophora carrionii AY857520<br />

Cladophialophora minourae AY251087<br />

Phialophora verrucosa AF050283<br />

Capronia semiimmersa AF050260<br />

Cladophialophora emmonsii AY857518<br />

Cladophialophora emmonsii AB109184<br />

Cladophialophora bantiana AY857517<br />

Cladophialophora sp. <strong>CBS</strong> 110551 AY857510<br />

Cladophialophora devriesii AB091212<br />

Cladophialophora devriesii AM114417<br />

Cladophialophora sp. <strong>CBS</strong> 102230 AY857508<br />

Cladophialophora sp. <strong>CBS</strong> 114326 AY<br />

Cladophialophora arxii AY857509<br />

Cladophialophora arxii AB109181<br />

Cladophialophora minourae AB091213<br />

Cladophialophora sp. AY781217<br />

Cladophialophora boppii AB109182<br />

Capronia fungicola AF050246<br />

83<br />

90<br />

94<br />

99<br />

98<br />

99<br />

80<br />

100<br />

100<br />

100<br />

Sorocybe resinae DAOM 239134, EU030275<br />

Capronia villosa AF050261<br />

Phaeococcomyces nigricans AY843154<br />

Ramichloridium cerophilum AF050286<br />

Heteroconium sp. AJ748260<br />

Phaeococcomyces catenatus AY843041<br />

Phaeococcomyces chersonesos AJ507323<br />

Metulocladosporiella musae DQ008138<br />

Metulocladosporiella musicola DQ008136<br />

Capronia acutiseta AF050241<br />

10 changes<br />

Fig. 6. Parsimony analysis of internal transcribed spacers sequences, demonstrating the position of Sorocybe resinae (shown in bold) among species of Capronia<br />

(Herpotrichiellaceae, Chaetothyriales) <strong>and</strong> its associated anamorph genera. One of 34 equally parsimonious trees (2 607 steps, CI = 0.301, RI = 0.506, RC = 0.153, HI = 0.699),<br />

with mid-point rooting. Bootstrap values above 70 % are shown at the relevant nodes; branches with thick lines occurred in all equally parsimonious trees.<br />

the same as one of his mutant forms of the creosote fungus,<br />

despite never having isolated such a dark spored form from any of<br />

his cultures. Parbery (1969) implied that the demonstrated ability<br />

of the creosote fungus to grow on a diversity of hydrocarbon-rich<br />

substrates favoured the thought that it would be able to grow on<br />

conifer resin. If cultures of the true Sorocybe resinae had been<br />

available, it is unlikely that this confusion would have persisted for<br />

so long. In vitro, the creosote fungus <strong>and</strong> the resin fungus are so<br />

different (Figs 1A, 2C) that it would difficult to defend the idea that<br />

they were mutants of the same fungus. <strong>The</strong>se differences in the<br />

cultures are reflected by the disparate phylogenetic affinities of<br />

what now are clearly demonstrated to be two different species.<br />

Unfortunately, the name Hormodendrum resinae has been<br />

misapplied to the creosote fungus, a species of economic<br />

importance. Also unfortunately, this species is the type of<br />

www.studiesinmycology.org<br />

Hormoconis, a generic name that the community concerned with<br />

this fungus has been slow to adapt to in the 30 years since its<br />

introduction. <strong>The</strong>re are several possible solutions to this problem.<br />

<strong>The</strong> conventional solution would be to apply names based strictly<br />

on the type specimens <strong>and</strong> accept Hormoconis as a synonym of<br />

Sorocybe, or to use it as a generic name for the mononematous<br />

synanamorph of the resin fungus. A new anamorph <strong>genus</strong> would<br />

then be described for the creosote fungus, making <strong>Cladosporium</strong><br />

avellaneum G.A. de Vries the basionym for its type. However, the<br />

resulting binomial would be unfamiliar to those concerned with the<br />

creosote fungus, <strong>and</strong> the earlier literature citing H. resinae would<br />

be misleading.<br />

A more parsimonious solution is possible. Article 14.9 of the<br />

International Code of Botanical Nomenclature (McNeil et al. 2006)<br />

allows for conservation of a name with a different type from that<br />

243


Seifert et al.<br />

designated by the authors. <strong>The</strong> name Hormodendrum resinae is<br />

not otherwise needed because the mononematous synanamorph<br />

of the resin fungus is rarely referred to by a Latin binomial, <strong>and</strong><br />

because Sorocybe resinae is based on a different type. <strong>The</strong>refore,<br />

a new type specimen could be proposed <strong>and</strong> conserved for<br />

Hormodendrum resinae Lindau, preferably the holotype of A.<br />

resinae (MELU 7130). This would make the anamorph-teleomorph<br />

connection unequivocal, maintain current species epithets <strong>and</strong><br />

taxonomic authorities, <strong>and</strong> ensure that most of the historical<br />

literature can be interpreted easily without the need to consult<br />

complicated nomenclators (Table 1). However, by perpetuating the<br />

use of the epithet “resinae”, this change would also perpetuate the<br />

misunderst<strong>and</strong>ing that resin is a possible substrate for the creosote<br />

fungus. In any case, the use of this epithet for the teleomorph of<br />

the creosote fungus, Amorphotheca resinae, is legitimate <strong>and</strong> valid,<br />

<strong>and</strong> unlikely ever to be changed.<br />

A third option would be an intermediate one. <strong>The</strong> application<br />

of the name <strong>Cladosporium</strong> avellaneum G.A. de Vries has never<br />

been in doubt, <strong>and</strong> it would be possible to conserve this species as<br />

the type of Hormoconis. This has the advantage of maintaining the<br />

familiar generic name Hormoconis, in combination with a species<br />

epithet that has been consistently applied. Furthermore, this solution<br />

would allow the confusion about the application <strong>and</strong> correct author<br />

citation around the epithet “resinae” for the anamorph of creosote<br />

fungus to recede.<br />

<strong>The</strong> second <strong>and</strong> third solutions require formal taxonomic<br />

proposals to be published in Taxon. We will argue the merits of<br />

these possible solutions at more length in that venue.<br />

<strong>The</strong> phylogenetic position of A. resinae raises additional<br />

taxonomic problems. This fungus typifies the monotypic family<br />

Amorphothecaceae, which has been considered incertae<br />

sedis since its description by Parbery (1969). Our phylogenetic<br />

analysis suggests that this family sits within the Myxotrichaceae.<br />

Amorphothecaceae (1969) is the older name, but Myxotrichaceae<br />

(1985) is well-entrenched in the mycological literature. As a<br />

consequence, the Myxotrichaceae are paraphyletic with respect<br />

to the Amorphothecaceae. <strong>The</strong> peridium of A. resinae, the only<br />

species presently placed in this family, lacks the thick-walled<br />

appendages that characterise most species of the Myxotrichaceae.<br />

Furthermore, the acropetal-blastic features of the anamorph of<br />

A. resinae differ from the thallic-arthric conidiogenesis of the<br />

other anamorphs associated with the Myxotrichaceae, principally<br />

Oidiodendron. <strong>The</strong>se morphological differences explain why the<br />

affinity of A. resinae with the Myxotrichaceae was not noted before.<br />

A formal proposal to conserve Myxotrichaceae as the name for this<br />

family might be prudent eventually, but this should await analysis of<br />

additional genes to confirm the phylogenetic relationship.<br />

Whether <strong>Cladosporium</strong> breviramosum, originally isolated from<br />

discoloured wallpaper, is actually a distinct species from A. resinae<br />

requires further study. It is clear that this species, if it is distinct,<br />

would be a member of Hormoconis rather than <strong>Cladosporium</strong>. Apart<br />

from the study of additional specimens, it might be fruitful to attempt<br />

to induce an Amorphotheca-like teleomorph in the two available<br />

cultures of C. breviramosum, <strong>and</strong> to compare the morphology with<br />

that of A. resinae. According to Parbery (1969), A. resinae includes<br />

both homothallic <strong>and</strong> heterothallic strains.<br />

Unfortunately, the phylogenetic affinities of Seifertia azaleae<br />

were not established with certainty in this study. Its closest relative<br />

in the LSU analysis is a sequence identified as Mycosphaerella<br />

mycopappi Funk & Dorworth (U43480, based on the apparent<br />

type culture ATCC 64711), but this sequence does not cluster<br />

with others representing the family Mycosphaerellaceae (data not<br />

shown). Similarly, the ITS sequences of the rhododendron fungus<br />

did not cluster with the many ITS sequences of Mycosphaerella<br />

available. Presently, it seems that Seifertia azaleae fungus is allied<br />

with the Dothideomycetes, but its precise affinities are uncertain. It<br />

is clear that this fungus should not be classified in Pycnostysanus<br />

(a taxonomic synonym of Sorocybe), <strong>and</strong> continued recognition of<br />

the monotypic <strong>genus</strong> Seifertia seems justified.<br />

ACKNOWLEDGEMENTS<br />

We are grateful to Sarah Hambleton <strong>and</strong> Marie Davey for providing ITS alignments<br />

that served as the basis for analyses in this paper. Walter Gams <strong>and</strong> Scott Redhead<br />

provided expert nomenclatural advice <strong>and</strong> helpful reviews of this manuscript.<br />

We thank Uwe Braun for proposing the third option for resolving the name of the<br />

creosote fungus, outlined in the text. <strong>The</strong> Canadian Collection of Fungal Cultures<br />

(DAOM) provided several of the strains for this study. <strong>The</strong> curators of DAOM <strong>and</strong> B<br />

kindly provided access to critical type specimens.<br />

REFERENCES<br />

Arx JA von (1973). Centraalbureau voor Schimmelcultures Baarn <strong>and</strong> Delft. Progress<br />

Reports 1972. Verh<strong>and</strong>elingen der Koninklijke Nederl<strong>and</strong>sche Akademie van<br />

Wetenschappen, Afdeeling Natuurkunde 61: 59–81.<br />

Bonorden HF (1851). H<strong>and</strong>buch der allgemeinen Mykologie. Stuttgart.<br />

Braun U, Crous PW, Dugan FM, Groenewald JZ, Hoog GS de (2003). Phylogeny<br />

<strong>and</strong> taxonomy of cladosporium-like hyphomycetes, including Davidiella gen.<br />

nov., the teleomorph of <strong>Cladosporium</strong> s. str. Mycological Progress 2: 3–18.<br />

Carmichael JW, Kendrick WB, Connors IL, Sigler L (1980). Genera of hyphomycetes.<br />

University of Alberta Press, Edmonton.<br />

Christensen CM, Kaufert FH, Schmitz H, Allison JL (1942). Hormodendrum resinae<br />

(Lindau), an inhabitant of wood impregnated with creosote <strong>and</strong> coal tar.<br />

American Journal of Botany 29: 552–558.<br />

Clements FC, Shear CL (1931). <strong>The</strong> genera of Fungi. H. W. Wilson, New York.<br />

Corda, ACJ (1839). Icones fungorum hucusque cognitorum, vol. 3. Prague.<br />

Davey ML, Currah RS (2007). A new species of Cladophialophora (hyphomycetes)<br />

from boreal <strong>and</strong> montane bryophytes. Mycological Research 111: 106–116.<br />

Ellis MB (1976). More Dematiaceous Hyphomycetes. Commonwealth Mycological<br />

Institute, Kew.<br />

Fries EM (1815). Observationes mycologicae, vol. 1, p. 216.<br />

Fries EM (1832). Systema Mycologicum, vol. 3(2). E. Moritz, Greifswald.<br />

Fries EM (1849). Summa vegatabilium Sc<strong>and</strong>inaviae,vol. 2: 468.<br />

Glawe DA, Hummell RL (2006). New North American host records for Seifertia<br />

azaleae. Pacific Northwest Fungi 1(5): 1–6.<br />

Hambleton S, Egger KN, Currah RS (1998). <strong>The</strong> <strong>genus</strong> Oidiodendron: species<br />

delimitation <strong>and</strong> phylogenetic relationships based on nuclear ribosomal DNA<br />

analysis. Mycologia 90: 854–869.<br />

Ho MH-M, Castañeda RF, Dugan FM, Jong, SC (1999). <strong>Cladosporium</strong> <strong>and</strong><br />

Cladophialophora in culture: descriptions <strong>and</strong> an exp<strong>and</strong>ed key. Mycotaxon<br />

72: 115–157.<br />

Hughes SJ (1958). Revisiones hyphomycetum aliquot cum appendice de nominibus<br />

rejiciendis. Canadian Journal of Botany 36: 727–836.<br />

Hughes SJ (1968). Strigopodia. Canadian Journal of Botany 46: 1099–1107.<br />

Katoh K, Misawa K, Kuma K, Miyata T (2002). MAFFT: a novel method for rapid<br />

multiple sequence alignment based on fast Fourier transform. Nucleic Acids<br />

Research 30: 3059–3066.<br />

Kendrick WB (1961). Hyphomycetes of conifer leaf litter. Hormodendrum<br />

staurophorum sp. nov. Canadian Journal of Botany 39: 833–835.<br />

Kornerup A, Wanscher JH (1989). Methuen H<strong>and</strong>book of Colour, 3 rd edn. Methuen,<br />

London.<br />

Lindau G (1904). Beiträge zur Pilzflora des Harzes. Verh<strong>and</strong>lungen des Botanischen<br />

Vereins der Provinz Br<strong>and</strong>enburg 45: 148–161 (‘1903’).<br />

Lindau G (1906). Dr. L. Rabenhorst’s Kryptogamen-Flora von Deutschl<strong>and</strong>,<br />

Oesterreich und der Schweiz. Zweite Auflage. Erster B<strong>and</strong>: Die Pilze<br />

Deutschl<strong>and</strong>s, Österreichs und der Schweiz. VIII. Abteilung: Fungi imperfecti:<br />

Hyphomycetes (erste Hälfte), Lief. 102: 641–704.<br />

Lindau G (1910). Dr. L. Rabenhorst’s Kryptogamen-Flora von Deutschl<strong>and</strong>,<br />

Oesterreich und der Schweiz. Zweite Auflage. Erster B<strong>and</strong>: Die Pilze<br />

Deutschl<strong>and</strong>s, Österreichs und der Schweiz. IX. Abteilung: Fungi imperfecti:<br />

Hyphomycetes (zweite Hälfte), Dematiaceae (Phaeophragmiae bis<br />

Phaeostaurosporae), Stilbaceae, Tuberculariaceae, sowie Nachträge,<br />

Nährpflanzenverzeichnis und Register. Leipzig.<br />

244


Hormoconis resinae <strong>and</strong> morphologically <strong>similar</strong> taxa<br />

Marsden DH (1954). Studies of the creosote fungus, Hormodendrum resinae.<br />

Mycologia 46: 161–183.<br />

Mason EW, Ellis MB (1953). British species of Periconia. Mycological Papers 56:<br />

1–127.<br />

McNeill J, Barrie FR, Burdet HM, Demoulin V, Hawksworth DL, Marhold K,<br />

Nicolson DH, Prado J, Silva PC, Skog JE, Wiersema JH, Turl<strong>and</strong> NJ (eds)<br />

(2007). International Code of Botanical Nomenclature (Vienna Code). Regnum<br />

Vegetabile 146: 1–568.<br />

Parbery DG (1969). Amorphotheca resinae gen. nov., sp. nov.: <strong>The</strong> perfect state of<br />

<strong>Cladosporium</strong> resinae. Australian Journal of Botany 17: 331–357.<br />

Partridge EC, Morgan-Jones G (2002). Notes on hyphomycetes, LXXXVIII. New<br />

genera in which to classify Alysidium resinae <strong>and</strong> Pycnostysanus azaleae, with<br />

a consideration of Sorocybe. Mycotaxon 83: 335–352.<br />

Persoon CH (1822). Mycologia europaea. Sectio Prima. Erlangae.<br />

Rehner SA, Samuels GJ (1995). Molecular systematics of the Hypocreales: a<br />

teleomorph gene phylogeny <strong>and</strong> the status of their anamorphs. Canadian<br />

Journal of Botany 73 (Suppl 1): S816–S823.<br />

Samson RA, Hoekstra ES, Frisvad JC (2004). Introduction to food- <strong>and</strong> airborne<br />

fungi. 7 th ed. Centraalbureau voor Schimmelcultures, Utrecht.<br />

Stafleu FA, Frans A, Mennega EA (1995). Taxonomic Literature. A selective guide<br />

to botanical publications <strong>and</strong> collections with dates, commentaries <strong>and</strong> types.<br />

Suppl. III (Br - Ca). Regnum Vegetabile 132: 1–550.<br />

Swofford DL (2002). PAUP*: Phylogenetic analysis using parsimony (*<strong>and</strong> other<br />

methods), version 4. Sinauer Associates, Sunderl<strong>and</strong>, Massachusetts.<br />

Thom C (1930). <strong>The</strong> Penicillia. Williams & Wilkins, Baltimore.<br />

Vilgalys R, Hester M (1990). Rapid identification <strong>and</strong> mapping of enzymatically<br />

amplified ribosomal DNA from several Cryptococcus species. Journal of<br />

Bacteriology 172: 4238–4246.<br />

Vries GA de (1952). Contribution to the knowledge of the <strong>genus</strong> <strong>Cladosporium</strong> Link<br />

ex Fries. <strong>The</strong>sis, University of Utrecht.<br />

Vries GA de (1955). <strong>Cladosporium</strong> avellaneum de Vries, a synonym of Hormodendrum<br />

resinae Lindau. Antonie van Leeuwenhoek 21: 166–168.<br />

White TJ, Bruns T, Lee S, Taylor J (1990). Amplification <strong>and</strong> direct sequencing of<br />

fungal ribosomal RNA sequences for phylogenetics. In: PCR Protocols: a guide<br />

to methods <strong>and</strong> applications. (Innis MA, Gelf<strong>and</strong> DH, Sninsky JJ, White TJ,<br />

eds.) Academic Press, San Diego, CA: 315–322.<br />

www.studiesinmycology.org<br />

245


INDEX OF FUNGAL NAMES<br />

Arranged according to <strong>genus</strong> <strong>and</strong> then epithets.<br />

Synonymised names are italicised <strong>and</strong> page numbers of main entries are bolded. A superscript asterisk ( * ) points to pages with illustrations,<br />

a superscript c ( c ) to pages with a cladogram, a superscript t ( t ) to pages with a table <strong>and</strong> a superscript k ( k ) to pages with a key to genera or<br />

species.<br />

Acidomyces, 12<br />

Acidomyces richmondensis, 12<br />

Acladium biophilum, 85<br />

Acladium herbarum, 105<br />

Acrotheca acuta, 84<br />

Acrotheca cerophila, 72<br />

Acrotheca multispora, 84<br />

Acrothecium multisporum, 84<br />

Amorphotheca, 235, 242<br />

Amorphotheca resinae, 235–236 * , 238 t , 239–241, 242 c , 244<br />

Annulatascus triseptatus, 62 c , 64 c<br />

Antennaria cellaris, 75<br />

Anungitea, 54 k , 185, 203–204, 216<br />

Anungitea fragilis, 203–204<br />

Anungitea heterospora, 204<br />

Anungitea longicatenata, 54 k<br />

Anungitea rhabdospora, 204<br />

Anungitea uniseptata, 211<br />

Anungitopsis, 203–204, 216<br />

Anungitopsis amoena, 163, 204, 207<br />

Anungitopsis intermedia, 204, 209<br />

Anungitopsis speciosa, 186, 188 c , 190 t , 193 c , 203–204, 216<br />

Apiosporina, 205, 216<br />

Apiosporina collinsii, 189 c , 191 t , 194 c , 205<br />

Aporothielavia leptoderma, 36 c –37 c<br />

Ascochyta pisi, 36 c –37 c<br />

Ascocoryne cylichnium, 240 c<br />

Ascosphaera apis, 240 c<br />

Ascotaiwania hughesii, 62 c , 64 c<br />

Asperisporium, 31, 216<br />

Athelia epiphylla, 6 c –7 c , 36 c –37 c , 58, 62 c , 64 c , 188 c , 193 c<br />

Aureobasidium caulivorum, 158 t<br />

Batcheloromyces, 4, 12 k<br />

Batcheloromyces eucalypti, 2 t , 6 c –7 c , 12<br />

Batcheloromyces leucadendri, 2 t , 6 c –7 c , 12–13 *<br />

Batcheloromyces proteae, 2 t , 6 c –7 c , 12, 63 c , 65 c , 162 c<br />

Bispora, 52 k<br />

Bisporella citrina, 36 c –37 c , 240 c<br />

Blumeria graminis f. sp. bromi, 36 c –37 c<br />

Botryosphaeria ribis , 97 c , 240 c<br />

Botryosphaeria stevensii , 97 c<br />

Botrytis epichloës, 89<br />

Byssoascus striatisporus, 240 c , 242 c<br />

Cadophora elatum, 46<br />

Calicium, 236<br />

Calicium viride, 240 c<br />

Calycina herbarum, 242 c<br />

Capnobotryella, 11 k –12, 17<br />

Capnobotryella renispora, 2 t , 5, 6 c –7 c , 12<br />

Capnodium citri, 240 c<br />

Capnodium coffeae, 6 c –7 c<br />

Capnodium salicinum, 2 t , 6 c –7 c<br />

Capronia, 55, 92, 185, 203, 216, 241–243<br />

Capronia acutiseta, 188 c , 193 c , 243 c<br />

Capronia coronata, 62 c , 64 c , 243 c<br />

Capronia dactylotricha, 243 c<br />

Capronia epimyces, 243 c<br />

Capronia fungicola, 243 c<br />

Capronia hanliniana, 185, 212<br />

Capronia hystrioides, 185, 212<br />

Capronia mansonii, 6 c –7 c , 62 c , 64 c , 188 c , 193 c , 240 c , 243 c<br />

Capronia munkii, 188 c , 193 c , 243 c<br />

Capronia nigerrima, 243 c<br />

Capronia parasitica, 243 c<br />

Capronia pilosella, 188 c , 193 c , 243 c<br />

Capronia pulcherrima, 62 c , 64 c , 243 c<br />

Capronia semiimmersa, 243 c<br />

Capronia villosa, 241, 243 c<br />

Caproventuria, 185, 216, 205, 230<br />

Caproventuria hanliniana, 189 c , 194 c , 205, 212<br />

Caproventuria hystrioides, 185, 212<br />

Carpoligna, 92<br />

Carpoligna pleurothecii, 62 c , 64 c , 83<br />

Castanedaea, 52 k<br />

Catenulostroma, 12 k , 13–14, 164<br />

Catenulostroma abietis, 2 t , 6 c –7 c , 14 k –15, 17, 63 c , 65 c<br />

Catenulostroma castellanii, 2 t , 6 c –7 c , 36 c –37 c<br />

Catenulostroma chromoblastomycosum, 2 t , 6 c –7 c , 14 k –15 * , 16,<br />

36 c –37 c<br />

Catenulostroma elginense, 2 t , 6 c –7 c , 15 k –16<br />

Catenulostroma excentricum, 10, 14 k , 16<br />

Catenulostroma germanicum, 2 t , 6 c –7 c , 14 k , 16 * –17, 36 c –37 c<br />

Catenulostroma macowanii, 2 t , 6 c –7 c , 15 k , 17, 63 c , 65 c , 162 c<br />

Catenulostroma microsporum, 2 t , 6 c –7 c , 10, 15 k , 17<br />

Catenulostroma protearum, 14–15 k , 17<br />

Catenulostroma sp., 2 t<br />

Ceramothyrium carniolicum, 188 c , 193 c , 240 c<br />

Cercospora, 1, 31, 164<br />

Cercospora apii , 97 c<br />

Cercospora beticola , 97 c , 111 c<br />

Cercospora cruenta, 162 c<br />

Cercospora epicoccoides, 11<br />

Cercospora eucalypti, 26<br />

Cercospora penicillata, 1<br />

Cercospora zebrina, 162 c<br />

Cercosporella centaureicola, 2 t , 6 c –7 c , 63 c , 65 c<br />

Cercostigmina, 30–31<br />

Chaetomium globosum, 36 c –37 c<br />

Chaetomium homopilatum, 36 c –37 c<br />

Chloridium apiculatum, 68<br />

Chloridium indicum, 68, 72–73<br />

Chloridium minus, 76<br />

Chloridium musae, 57, 72<br />

Chloridium schulzerii, 84<br />

Cibiessia, 11 k , 17, 24–25, 30<br />

Cibiessia dimorphospora, 2 t , 6 c –7 c , 17, 63 c , 65 c<br />

Cibiessia minutispora, 2 t , 6 c –7 c<br />

Cibiessia nontingens, 2 t , 6 c –7 c<br />

Cistella acuum, 214<br />

246


Cladophialophora, 52 k , 54 k –55, 167, 185–188, 198–200, 204, 216,<br />

219–220, 226–228, 230–232<br />

Cladophialophora ajelloi, 231<br />

Cladophialophora arxii, 243 c<br />

Cladophialophora australiensis, 186, 187–188 c , 190 t , 193 c –194 *<br />

Cladophialophora bantiana, 187, 231, 243 c<br />

Cladophialophora boppii, 231, 243 c<br />

Cladophialophora brevicatenata, 185, 212<br />

Cladophialophora carrionii, 185, 187–188 c , 193 c , 219, 220 t –221 t ,<br />

222–223, 224 c –226 c , 227–228 * , 229–230 * , 231–232, 243 c<br />

Cladophialophora cf. carrionii, 232<br />

Cladophialophora chaetospira, 54 k , 187–188 c , 190 t , 193 c , 195 *<br />

Cladophialophora devriesii, 231, 243 c<br />

Cladophialophora emmonsii, 243 c<br />

Cladophialophora hachijoensis, 185<br />

Cladophialophora hostae, 187, 188 c –189, 190 t , 193 c , 195 * –196 * ,<br />

199, 204<br />

Cladophialophora humicola, 188 c –189, 190 t , 193 c , 196 *<br />

Cladophialophora minourae, 243 c<br />

Cladophialophora potulentorum, 188 c , 190 t , 193 c , 197 * –198<br />

Cladophialophora proteae, 188 c , 190 t , 193 c , 198 *<br />

Cladophialophora scillae, 188 c , 190 t , 193 c , 198–199 * , 204<br />

Cladophialophora sp., 223, 225, 243 c<br />

Cladophialophora sylvestris, 188 c , 190 t , 193 c , 200 *<br />

Cladophialophora yegresii, 185, 221 t , 224 c , 226 c , 227–228, 229 * –<br />

230 * , 231–232<br />

Cladoriella, 55<br />

Cladoriella eucalypti, 34 t , 36 c –37 c<br />

<strong>Cladosporium</strong>, 8, 21, 30, 33, 35, 38–39, 41, 44, 47, 49, 52 k , 55,<br />

57, 72, 95–96, 103, 105–106 * , 107, 113, 115–116, 137, 142,<br />

153–155, 157, 161–169, 173, 176, 181, 209, 227, 230, 237–<br />

238, 242, 244<br />

“<strong>Cladosporium</strong>” adianticola, 190 t , 193 c<br />

<strong>Cladosporium</strong> alneum, 116–117<br />

<strong>Cladosporium</strong> alopecuri, 142<br />

<strong>Cladosporium</strong> amoenum, 207<br />

<strong>Cladosporium</strong> antarcticum, 108 t , 111 c –112 c , 114 k –115, 116 * –117 * ,<br />

154–155<br />

<strong>Cladosporium</strong> araguatum, 43<br />

<strong>Cladosporium</strong> argillaceum, 50<br />

<strong>Cladosporium</strong> arthoniae, 41, 116<br />

<strong>Cladosporium</strong> arthrinioides, 140<br />

<strong>Cladosporium</strong> arthropodii, 142<br />

<strong>Cladosporium</strong> asterinae, 55, 103<br />

<strong>Cladosporium</strong> avellaneum, 235, 237–238 t , 242–244<br />

<strong>Cladosporium</strong> breviramosum, 241–242 c , 244<br />

<strong>Cladosporium</strong> bruhnei, 2 t , 6 c –7 c , 36 c –37 c , 63 c , 65 c , 97 c , 107–108 t ,<br />

110, 111 c –112 c , 114 k , 118 * –120 * , 145, 154–155, 158 t , 162 c –165 c ,<br />

168<br />

<strong>Cladosporium</strong> brunneum, 133<br />

<strong>Cladosporium</strong> carpophilum, 205<br />

<strong>Cladosporium</strong> castellanii, 43, 44<br />

<strong>Cladosporium</strong> cellare, 75<br />

<strong>Cladosporium</strong> cerophilum, 72<br />

<strong>Cladosporium</strong> chlorocephalum, 95–96, 103<br />

<strong>Cladosporium</strong> cladosporioides, 2 t , 6 c –7 c , 36 c –37 c , 63 c , 65 c , 108 t ,<br />

111 c –112 c , 113 k , 140, 147, 150, 153, 158 t , 162 c –165 c , 168, 174<br />

<strong>Cladosporium</strong> colocasiae, 162 c<br />

<strong>Cladosporium</strong> cubense, 19–20<br />

<strong>Cladosporium</strong> dominicanum, 158 t , 162 c –165 c , 166 * , 168 k –169,<br />

170 * , 175<br />

<strong>Cladosporium</strong> elatum, 46<br />

<strong>Cladosporium</strong> ferrugineum, 20–21<br />

<strong>Cladosporium</strong> fulvum, 55<br />

<strong>Cladosporium</strong> fusiforme, 158 t , 162 c –165 c , 166 * , 168 k –169, 171 *<br />

<strong>Cladosporium</strong> gracile, 133<br />

<strong>Cladosporium</strong> graminum, 133<br />

<strong>Cladosporium</strong> halotolerans, 158 t , 162 c –165 c , 166 * –167, 168 k –169,<br />

172 * –173, 175<br />

<strong>Cladosporium</strong> helicosporum, 18<br />

<strong>Cladosporium</strong> hemileiae, 55<br />

<strong>Cladosporium</strong> herbaroides, 108 t , 111 c –112 c , 114 k , 120, 121 * –123 * ,<br />

154<br />

<strong>Cladosporium</strong> herbarum, 8, 33, 96–97 c , 101, 105, 107–108 t , 110,<br />

111 c –112 c , 113–114 k , 118, 120, 122, 124, 125 * –127 * , 133, 142,<br />

153–155, 158 t , 162 c , 164–166, 168, 181<br />

<strong>Cladosporium</strong> herbarum var. macrocarpum, 105, 129<br />

<strong>Cladosporium</strong> hordei, 118<br />

<strong>Cladosporium</strong> hypophyllum, 140<br />

<strong>Cladosporium</strong> iridis, 36 c –37 c , 97 c , 109 t , 111 c –112 c , 114 k , 125, 127–<br />

128 * , 135, 137, 153<br />

<strong>Cladosporium</strong> langeronii, 159 t , 162 c –165 c , 166 * –167, 168 k , 173–<br />

174 *<br />

<strong>Cladosporium</strong> lichenicola, 116<br />

<strong>Cladosporium</strong> licheniphilum, 116<br />

<strong>Cladosporium</strong> lichenum, 116<br />

<strong>Cladosporium</strong> macrocarpum, 97 c , 105, 107, 109 t –110, 111 c –112 c ,<br />

114 k , 122, 129 * –130 * , 131 * –133, 153–155<br />

<strong>Cladosporium</strong> magnusianum, 133, 135<br />

<strong>Cladosporium</strong> musae, 55<br />

<strong>Cladosporium</strong> nigrellum, 209<br />

<strong>Cladosporium</strong> olivaceum, 237<br />

<strong>Cladosporium</strong> ossifragi, 97 c , 109 t –110, 111 c –112 c , 114 k , 133 * –134 * ,<br />

135, 154<br />

<strong>Cladosporium</strong> oxysporum, 159 t , 162 c –165 c , 168<br />

<strong>Cladosporium</strong> paeoniae, 95–96, 101, 103<br />

<strong>Cladosporium</strong> paeoniae var. paeoniaeanomalae, 95–96<br />

<strong>Cladosporium</strong> pseudiridis, 109 t , 111 c –112 c , 114 k , 135, 136 * –137 *<br />

<strong>Cladosporium</strong> psoraleae, 116–117<br />

<strong>Cladosporium</strong> psychrotolerans, 159 t , 162 c –165 c , 166 * –167, 168 k ,<br />

175–176 *<br />

<strong>Cladosporium</strong> ramotenellum, 107, 109 t , 111 c –112 c , 115 k , 137–138 * ,<br />

139 * –140, 150, 154, 159 t , 163 c<br />

<strong>Cladosporium</strong> resinae, 235, 237–238 t<br />

<strong>Cladosporium</strong> rigidophorum, 21, 22<br />

<strong>Cladosporium</strong> salinae, 159 t , 161, 162 c –165 c , 166 * , 168 k , 175,<br />

176–177 *<br />

<strong>Cladosporium</strong> scillae, 198<br />

<strong>Cladosporium</strong> sinuosum, 109 t , 111 c –112 c , 114 k , 140 * –141 * , 142<br />

<strong>Cladosporium</strong> sp., 109 t , 111 c , 149, 159 t , 163 c , 165 c<br />

<strong>Cladosporium</strong> sphaerospermum, 2 t , 6 c –7 c , 63 c , 65 c , 153, 157, 159 t ,<br />

162 c –165 c , 166 * , 167–168 k , 169, 173–175, 177–178 * , 181<br />

<strong>Cladosporium</strong> spinulosum, 109 t , 111 c –112 c , 114 k , 142 * , 145, 154–<br />

155, 160 t , 162 c –165 c , 166 * , 168 k , 179 * , 180–181<br />

<strong>Cladosporium</strong> staurophorum, 55<br />

<strong>Cladosporium</strong> strumelloideum, 23<br />

<strong>Cladosporium</strong> subinflatum, 109 t , 111 c –112 c , 114 k , 143 * –144 * , 145,<br />

154, 160 t , 163 c –165 c<br />

<strong>Cladosporium</strong> subnodosum, 150<br />

<strong>Cladosporium</strong> subtilissimum, 109 t , 111 c –112 c , 114 k , 145 * –146 * ,<br />

147 * , 149–150, 154–155<br />

<strong>Cladosporium</strong> tenellum, 97 c , 109 t , 111 c –112 c , 115 k , 137, 140, 148 * –<br />

149 * , 150, 153–154<br />

247


<strong>Cladosporium</strong> tenuissimum, 160 t , 162 c –163 c , 168<br />

<strong>Cladosporium</strong> uredinicola, 2 t , 6 c –7 c , 36 c –37 c , 63 c , 65 c , 116, 162 c<br />

<strong>Cladosporium</strong> variabile, 97 c , 110 t , 111 c –112 c , 114 k , 131, 150 * –152 * ,<br />

153–154<br />

<strong>Cladosporium</strong> velox, 160 t , 162 c –165 c , 166 * , 168 k , 180 * –181<br />

Clathrosporium intricatum, 189 c , 193 c<br />

Coccodinium, 28<br />

Coccodinium bartschii, 2 t , 4–5, 6 c –7 c , 28, 29 *<br />

Cochliobolus heliconiae, 240 c<br />

Colletogloeopsis, 24<br />

Colletogloeopsis blakelyi, 26<br />

Colletogloeopsis considenianae, 26<br />

Colletogloeopsis dimorpha, 26<br />

Colletogloeopsis gauchensis, 26<br />

Colletogloeopsis molleriana, 10<br />

Colletogloeopsis nubilosum, 10<br />

Colletogloeopsis stellenboschiana, 26<br />

Colletogloeopsis zuluensis, 26<br />

Colletogloeum nubilosum, 10<br />

Coniella, 25<br />

Conioscypha lignicola, 62 c , 64 c<br />

Conioscyphascus varius, 62 c , 64 c<br />

Coniosporium, 216<br />

Coniothecium macowanii, 17<br />

Coniothecium punctiforme, 17<br />

Coniothyrium palmarum, 34 t , 36 c –37 c<br />

Coniothyrium zuluense, 26<br />

Coremium, 236, 239<br />

Cryptadelphia groenendalensis, 62 c , 64 c<br />

Cryptadelphia polyseptata, 62 c , 64 c<br />

Cryptococcus neoformans, 232<br />

Cudonia lutea, 240 c<br />

Cylindrosympodium, 204<br />

Cylindrosympodium lauri, 189 c –190 t , 194 c , 204 *<br />

Cylindrosympodium variabile, 205<br />

Cyphellophora, 201<br />

Cyphellophora fusarioides, 201<br />

Cyphellophora guyanensis, 201<br />

Cyphellophora hylomeconis, 188 c , 190 t , 193 c , 200 * –201<br />

Cyphellophora laciniata, 188 c , 190 t , 193 c , 201<br />

Cyphellophora pluriseptata, 201<br />

Cyphellophora suttonii, 201<br />

Cyphellophora vermispora, 201<br />

Davidiella, 5 k , 8, 30, 33, 35, 44, 52 k , 92, 96, 106, 115, 153, 155,<br />

163, 166, 173, 230<br />

Davidiella allicina, 2 t , 108 t , 112 c , 115 k , 118, 119 * –120 *<br />

Davidiella macrocarpa, 109 t , 112 c , 115 k , 129–130 * , 132 * –133<br />

Davidiella macrospora, 109 t , 112 c , 125, 128<br />

Davidiella sp., 108 t , 110 t , 111 c –112 c , 155<br />

Davidiella tassiana, 8, 108 t , 112 c , 115 k , 120, 122, 124 * –125, 126 * ,<br />

158 t , 163 c –165 c , 166, 168<br />

Davidiella variabile, 110 t , 112 c , 115 k , 151 * , 152–153<br />

Dematium epiphyllum var. chionanthi, 122, 124<br />

Dematium graminum, 129, 132<br />

Dematium herbarum, 8, 105, 122<br />

Dematium herbarum var. brassicae, 129, 132<br />

Dematium nigrum, 238 t –239<br />

Dematium vulgare var. foliorum, 129, 132<br />

Dematium vulgare var. typharum, 129, 132<br />

Dendryphion resinae, 238 t<br />

Denticularia, 31<br />

Devriesia, 11 k , 17, 43, 54 k –55, 164, 227, 230<br />

Devriesia acadiensis, 34 t , 36 c –37 c<br />

Devriesia americana, 34 t , 36 c –37 c , 42–43 *<br />

Devriesia shelburniensis, 34 t , 36 c –37 c , 43<br />

Devriesia staurophora, 5, 6 c –7 c , 17, 36 c –37 c , 63 c , 65 c , 162 c<br />

Devriesia thermodurans, 34 t , 36 c –37 c<br />

Dichocladosporium, 53 k , 55, 96, 103<br />

Dichocladosporium chlorocephalum, 36 c –37 c , 96, 97 c , 98 t* –99 * ,<br />

100 * –101, 102 *<br />

Didymella bryoniae, 36 c –37 c<br />

Didymella cucurbitacearum, 36 c –37 c<br />

Didymellina macrospora, 126<br />

Digitopodium, 52 k , 55<br />

Diplococcium, 39<br />

Discosphaerina fagi, 162 c , 240 c<br />

Dissoconium, 5 k , 28, 61<br />

Dissoconium aciculare, 2 t , 6 c –7 c , 28, 63 c , 65 c<br />

Dissoconium commune, 2 t , 28<br />

Dissoconium dekkeri, 2 t<br />

Distocercospora, 31<br />

Dothidea insculpta, 162 c<br />

Dothidea ribesia, 162 c<br />

Dothistroma, 4<br />

Exophiala, 75, 92, 185, 201, 203, 216, 231<br />

Exophiala attenuata, 243 c<br />

Exophiala crusticola, 243 c<br />

Exophiala dermatitidis, 6 c –7 c , 62 c , 64 c , 188 c , 193 c , 201, 243 c<br />

Exophiala eucalyptorum, 188 c , 190 t , 193 c , 203 *<br />

Exophiala heteromorpha, 243 c<br />

Exophiala jeanselmei, 6 c –7 c , 62 c , 64 c , 188 c , 193 c<br />

Exophiala lecaniicorni, 243 c<br />

Exophiala mesophila, 201, 243 c<br />

Exophiala nishimurae, 243 c<br />

Exophiala oligosperma, 6 c –7 c , 188 c , 193 c , 243 c<br />

Exophiala phaeomuriformis, 201<br />

Exophiala pisciphila, 188 c , 193 c , 243 c<br />

Exophiala sp. 1, 188 c , 190 t , 193 c , 201–202 *<br />

Exophiala sp. 2, 188 c , 190 t , 193 c , 201–202 *<br />

Exophiala sp. 3, 188 c , 190 t , 193 c<br />

Exophiala spinifera, 243 c<br />

Fasciatispora petrakii, 37 c , 188 c , 193 c<br />

Fonsecaea, 231<br />

Fonsecaea monophora, 231<br />

Fonsecaea pedrosoi, 6 c –7 c , 62 c , 64 c , 188 c , 193 c , 219, 232, 240 c<br />

Friedmanniomyces, 11 k –12, 24<br />

Fumagospora capnodioides, 2 t<br />

Fusarium, 231<br />

Fusicladium, 17, 52 k , 54 k –55, 91, 96, 103, 163, 176, 185, 188–189,<br />

197–199, 203–204, 205, 211, 216<br />

Fusicladium africanum, 189 c –190 t , 194 c , 205–206 * , 211<br />

Fusicladium amoenum, 162 c , 189 c –190 t , 194 c , 206 * –207<br />

Fusicladium brevicatenatum, 212<br />

Fusicladium carpophilum, 189 c –190 t , 194 c<br />

Fusicladium caruanianum, 208<br />

Fusicladium catenosporum, 189 c , 191 t , 194 c<br />

Fusicladium convolvularum, 189 c , 191 t , 194 c , 207 * –208<br />

Fusicladium effusum, 161, 162 c –163, 189 c , 191 t , 194 c<br />

Fusicladium fagi, 189 c , 191 t , 194 c , 208 * –209<br />

“Fusicladium hyphopodioides”, 215<br />

Fusicladium intermedium, 189 c , 191 t , 194 c , 209 *<br />

Fusicladium m<strong>and</strong>shuricum, 189 c , 191 t , 194 c , 210<br />

Fusicladium matsushimae, 209<br />

Fusicladium oleagineum, 189 c , 191 t , 194 c<br />

248


Fusicladium phillyreae, 189 c , 191 t , 194 c<br />

Fusicladium pini, 189 c , 191 t , 194 c , 210 * , 211–212<br />

Fusicladium pomi, 189 c , 191 t , 194 c<br />

Fusicladium radiosum, 189 c , 191 t , 194 c<br />

Fusicladium ramoconidii, 189 c , 191 t , 194 c , 211 * –212 *<br />

Fusicladium rhodense, 189 c , 191 t , 194 c , 212–213 *<br />

Fusicladium scillae, 198–199<br />

Fusicladium sp., 212<br />

Fusicladosporium, 205<br />

Geomyces asperulatus, 242 c<br />

Geomyces pannorum, 242 c<br />

Glyphium elatum, 188 c , 193 c , 240 c<br />

Golovinomyces cichoracearum, 240 c –241<br />

Gomphinaria, 84<br />

Gomphinaria amoena, 84<br />

Graphiopsis chlorocephala, 96<br />

Hansfordia biophila, 85<br />

Hansfordia torvi, 85<br />

Haplographium chlorocephalum, 96<br />

Haplographium chlorocephalum var. ovalisporum, 96<br />

Haplotrichum, 54 k<br />

Helminthosporium, 239<br />

Helminthosporium variabile, 150<br />

Hendersonia gr<strong>and</strong>ispora, 11<br />

Heteroconium, 54 k , 187, 241<br />

Heteroconium chaetospira, 187, 216<br />

Heteroconium citharexyli, 187<br />

Heteroconium sp., 243 c<br />

Heterosporium, 33, 142, 153<br />

Heterosporium iridis, 125<br />

Heterosporium magnusianum, 133, 135<br />

Heterosporium ossifragi, 133, 135<br />

Heterosporium variabile, 150<br />

Heyderia abietis, 242 c<br />

Hormiactis, 51 k<br />

Hormoconis, 54 k , 235, 238, 243–244<br />

Hormoconis resinae, 34 t , 36 c –37 c , 235, 238 t , 240 c , 243<br />

“Hormodendron”, 237<br />

Hormodendrum, 237<br />

“Hormodendrum” elatum, 47<br />

Hormodendrum hordei, 118, 120<br />

Hormodendrum langeronii, 167, 173–174<br />

Hormodendrum olivaceum, 237<br />

Hormodendrum resinae, 54 k , 235, 237–238 t , 239–240 * , 241–244<br />

Hormodendrum viride, 237<br />

Hortaea, 12 k , 17<br />

Hortaea werneckii, 2 t , 6 c –7 c , 17<br />

Hyalodendriella, 46, 52 k<br />

Hyalodendriella betulae, 34 t , 36 c –37 c , 46–47 *<br />

Hyalodendron, 46, 51 k<br />

Iodosphaeria aquatica, 240 c<br />

Kirramyces, 24<br />

Kirramyces destructans, 26<br />

Kirramyces epicoccoides, 11<br />

Kirramyces eucalypti, 26<br />

Kumbhamaya, 201<br />

Lacazia loboi, 167<br />

Lecanosticta, 24<br />

Leptodontidium elatius, 242 c<br />

Leptospora rubella, 36 c –37 c<br />

Letendraea helminthicola, 240 c<br />

Loboa loboi, 159 t<br />

Lojkania enalia, 240 c<br />

Lophodermium pinastri, 240 c<br />

Lylea, 51 k<br />

Madurella mycetomatis, 232<br />

Magnaporthe grisea, 62 c , 64 c<br />

Melanchlenus eumetabolus, 243 c<br />

Melanchlenus oligospermus, 243 c<br />

Melanelia exasperatula, 240 c<br />

Metacoleroa, 205, 216<br />

Metacoleroa dickiei, 62 c , 64 c , 189 c , 194 c , 205<br />

Metulocladosporiella, 18, 51, 53 k , 55, 103<br />

Metulocladosporiella musae, 103, 188 c , 193 c , 243 c<br />

Metulocladosporiella musicola, 188 c , 193 c , 243 c<br />

Mollisia cinerea, 6 c –7 c<br />

Mycosphaerella, 1, 5 k , 8, 28, 30, 31–33, 45, 51 k , 55, 91, 96, 105,<br />

115, 155, 163, 164, 166, 205, 241–242, 244<br />

Mycosphaerella africana, 9, 36 c –37 c<br />

Mycosphaerella alchemillicola, 34 t<br />

Mycosphaerella alistairii, 8–9<br />

Mycosphaerella ambiphylla, 10<br />

Mycosphaerella associata, 8–9<br />

Mycosphaerella aurantia , 97 c<br />

Mycosphaerella bellula, 8, 10<br />

“Mycosphaerella” communis, 2 t , 6 c –7 c , 28, 63 c , 65 c<br />

Mycosphaerella cryptica, 8, 10<br />

Mycosphaerella dendritica, 8, 10<br />

Mycosphaerella endophytica, 63 c , 65 c<br />

Mycosphaerella eucalypti, 8<br />

Mycosphaerella excentrica, 8, 10<br />

Mycosphaerella fibrillosa, 10<br />

Mycosphaerella fimbriata, 8, 10<br />

Mycosphaerella flexuosa, 10<br />

Mycosphaerella gamsii, 10<br />

Mycosphaerella graminicola, 3 t , 6 c –7 c , 63 c , 65 c , 162 c<br />

Mycosphaerella gr<strong>and</strong>is, 10<br />

Mycosphaerella gregaria, 63 c , 65 c<br />

“Mycosphaerella” hyperici, 158 t , 163 c –165 c , 173<br />

Mycosphaerella intermedia, 63 c , 65 c<br />

Mycosphaerella irregulariramosa, 36 c –37 c<br />

Mycosphaerella jonkershoekensis, 8, 10<br />

Mycosphaerella juvenis, 10<br />

Mycosphaerella latebrosa, 162 c<br />

“Mycosphaerella” lateralis, 2 t , 6 c –7 c , 28, 36 c –37 c , 63 c , 65 c , 162 c<br />

Mycosphaerella macrospora, 126<br />

Mycosphaerella madeirae, 63 c , 65 c<br />

Mycosphaerella marksii, 36 c –37 c , 63 c , 65 c<br />

Mycosphaerella maxii, 8, 10<br />

Mycosphaerella mexicana, 8, 10<br />

Mycosphaerella microspora, 10<br />

Mycosphaerella molleriana, 10<br />

“Mycosphaerella mycopappi”, 240 c –241, 244<br />

Mycosphaerella nubilosa, 8, 10<br />

Mycosphaerella ohnowa, 10<br />

Mycosphaerella parkii, 63 c , 65 c<br />

Mycosphaerella parkiiaffinis, 10<br />

Mycosphaerella parva, 10<br />

Mycosphaerella perpendicularis, 10<br />

Mycosphaerella pluritubularis, 10<br />

Mycosphaerella pseudafricana, 11<br />

Mycosphaerella pseudocryptica, 8, 11<br />

Mycosphaerella pseudosuberosa, 8, 11<br />

Mycosphaerella punctiformis, 6 c –7 c , 28, 36 c –37 c , 63 c , 65 c , 97 c ,<br />

166<br />

249


Mycosphaerella quasicercospora, 11<br />

Mycosphaerella readeriellophora, 11<br />

Mycosphaerella secundaria, 11<br />

Mycosphaerella stigminaplatani, 30<br />

Mycosphaerella stramenticola, 11<br />

Mycosphaerella suberosa, 8, 11<br />

Mycosphaerella suttonii, 11<br />

Mycosphaerella tassiana, 122, 166<br />

Mycosphaerella toledana, 11<br />

Mycosphaerella tulasnei, 132<br />

Mycosphaerella vespa, 10<br />

Mycosphaerella walkeri, 63 c , 65 c<br />

Mycovellosiella, 31, 53 k<br />

Myrmecridium, 61 k , 84, 92<br />

Myrmecridium flexuosum, 59 t , 62 c , 64 c , 84–85, 86 *<br />

Myrmecridium schulzeri, 59 t , 62 c , 64 c , 68 * , 84–85 *<br />

Myrmecridium schulzeri var. schulzeri, 84<br />

Myrmecridium schulzeri var. tritici, 84<br />

Myxotrichum, 242<br />

Myxotrichum arcticum, 242 c<br />

Myxotrichum cancellatum, 242 c<br />

Myxotrichum carminoparum, 242 c<br />

Myxotrichum chartarum, 242 c<br />

Myxotrichum deflexum, 36 c –37 c , 242 c<br />

Myxotrichum resinae, 239<br />

Myxotrichum stipitatum, 242 c<br />

Napicladium ossifragi, 133, 135<br />

Neofabraea alba, 36 c –37 c , 240 c , 242 c<br />

Neofabraea krawtzewii, 242 c<br />

Neofabraea malicorticis, 36 c –37 c , 189 c , 193 c , 242 c<br />

Neofabraea perennans, 242 c<br />

Neoovularia, 87<br />

Nothostrasseria, 12 k<br />

Nothostrasseria dendritica, 2 t , 6 c –7 c , 10<br />

Ochrocladosporium, 46, 54 k –55<br />

Ochrocladosporium elatum, 34 t , 36 c –37 c , 46, 48 * –49<br />

Ochrocladosporium frigidarii, 34 t , 36 c –37 c , 47–49 *<br />

Oidiodendron, 241–242, 244<br />

Oidiodendron cerealis, 242 c<br />

Oidiodendron chlamydosporicum, 242 c<br />

Oidiodendron citrinum, 242 c<br />

Oidiodendron flavum, 242 c<br />

Oidiodendron griseum, 242 c<br />

Oidiodendron maius, 242 c<br />

Oidiodendron myxotrichoides, 242 c<br />

Oidiodendron periconioides, 242 c<br />

Oidiodendron pilicola, 242 c<br />

Oidiodendron rhodogenum, 242 c<br />

Oidiodendron setiferum, 242 c<br />

Oidiodendron tenuissimum, 242 c<br />

Oidiodendron truncatum, 242 c<br />

Ophiostoma stenoceras, 62 c , 64 c<br />

Paracercospora, 31<br />

Paracoccidioides loboi, 167<br />

Parahaplotrichum, 54 k<br />

Parapericoniella, 53 k , 55, 103<br />

Parapleurotheciopsis, 51, 53 k –54 k , 204<br />

Parapleurotheciopsis coccolobae, 51, 54 k<br />

Parapleurotheciopsis inaequiseptata, 34 t –35, 37 c , 51<br />

Passalora, 28, 31, 44–45, 53 k , 216<br />

Passalora arachidicola , 97 c<br />

Passalora daleae, 34 t , 36 c –37 c<br />

Passalora dodonaeae, 162 c<br />

Passalora eucalypti, 36 c –37 c<br />

Passalora fulva, 36 c –37 c , 55, 97 c , 162 c<br />

Passalora perplexa, 30<br />

“Passalora” zambiae, 2 t , 6 c –7 c , 28<br />

Paullicorticium ansatum, 6 c –7 c , 36 c –37 c , 58, 62 c , 64 c , 188 c , 193 c<br />

Penicillium, 237<br />

Penicillium olivaceum, 237<br />

Penidiella, 12 k , 17–18, 21–22, 24, 53 k<br />

Penidiella columbiana, 2 t , 6 c –7 c , 17–18 k , 19 * –20 * , 36 c –37 c<br />

Penidiella cubensis, 18 k , 19–21 *<br />

Penidiella nect<strong>and</strong>rae, 3 t , 18 k , 20–21 * , 36 c –37 c<br />

Penidiella rigidophora, 3 t , 18 k , 21, 22 * –23 * , 36 c –37 c<br />

Penidiella strumelloidea, 3 t , 18 k , 23, 23 * –24 * , 36 c –37 c , 53 k<br />

Penidiella venezuelensis, 3 t , 18 k , 22, 24–25 * , 36 c –37 c<br />

Periconia chlorocephala, 95–96, 101, 103<br />

Periconia ellipsospora, 96<br />

Periconia velutina, 64<br />

Periconiella, 17, 53 k , 57–58, 61 k –62, 63, 91<br />

Periconiella arcuata, 59 t , 63 c , 65 c –66 * , 68 *<br />

Periconiella levispora, 59 t , 63 c , 65 c , 67 * –68<br />

Periconiella musae, 57, 72<br />

Periconiella papuana, 60, 80<br />

Periconiella velutina, 17, 57, 59 t –63 c , 64–65 c , 66 *<br />

Pertusaria mammosa, 240 c<br />

Phaeoblastophora, 52 k<br />

Phaeococcomyces catenatus, 188 c , 193 c , 243 c<br />

Phaeococcomyces chersonesos, 243 c<br />

Phaeococcomyces nigricans, 241, 243 c<br />

Phaeoisariopsis, 4, 31<br />

Phaeophleospora, 4, 24<br />

Phaeophleospora destructans, 26<br />

Phaeophleospora epicoccoides, 11<br />

Phaeophleospora eucalypti, 26<br />

Phaeophleospora toledana, 11, 25<br />

Phaeoramularia, 31, 53 k , 227<br />

“Phaeoramularia” hachijoensis, 43, 185, 214<br />

Phaeoseptoria eucalypti, 11<br />

Phaeoseptoria luzonensis, 11<br />

Phaeosphaeria avenaria, 36 c –37 c<br />

Phaeotheca triangularis, 3 t , 6 c –7 c<br />

Phaeothecoidea, 11 k<br />

Phaeothecoidea eucalypti, 3 t , 6 c –7 c<br />

Phaeotrichum benjaminii, 240 c<br />

Phialophora, 231<br />

Phialophora americana, 188 c , 193 c , 240 c<br />

Phialophora verrucosa, 243 c<br />

Phlogicylindrium, 216<br />

Phlogicylindrium eucalypti, 37 c , 188 c , 193 c , 214<br />

Phoma herbarum, 36 c –37 c<br />

Phyllactinia guttata, 132<br />

Phyllactinia sp., 110<br />

Piceomphale bulgarioides, 242 c<br />

Pidoplitchkoviella terricola, 188 c , 193 c<br />

Pilidiella, 25<br />

Plectosphaera eucalypti, 37 c<br />

Pleomassaria siparia, 240 c<br />

Pleospora herbarum, 240 c<br />

Pleuroascus nicholsonii, 240 c<br />

Pleurophoma sp., 3 t<br />

Pleurophragmium acutum, 84<br />

Pleurophragmium tritici, 84<br />

Pleurotheciopsis, 18, 51<br />

250


Ramichloridium epichloës, 60, 89<br />

Pleurothecium, 61 k , 83, 92<br />

243 c Rhinocladiella sp., 60 t , 62 c , 64 c<br />

Pleurothecium obovoideum, 59 t , 62 c , 64 c , 80, 83 * , 92<br />

Pollaccia, 185, 205<br />

Pollaccia m<strong>and</strong>shurica, 210<br />

Pollaccia sinensis, 210<br />

Polyscytalum, 51 k , 55 k , 185, 200, 214, 216<br />

Polyscytalum fecundissimum, 188 c , 191 t , 193 c , 200, 214 *<br />

Polyscytalum griseum, 200<br />

Prathigada, 31<br />

Protoventuria alpina, 186, 189 c , 191 t , 193 c<br />

Pseudeurotium bakeri, 242 c<br />

Pseudeurotium desertorum, 242 c<br />

Pseudeurotium ovale var. milkoi, 242 c<br />

Pseudeurotium ovale var. ovale, 242 c<br />

Pseudeurotium zonatum, 242 c<br />

Pseudocercospora, 4, 12, 30–31, 164<br />

Pseudocercospora angolensis, 162 c<br />

Pseudocercospora epispermogonia, 63 c , 65 c<br />

Pseudocercospora paraguayensis, 36 c –37 c<br />

Pseudocercospora protearum, 162 c<br />

Pseudocercosporidium, 31<br />

Ramichloridium fasciculatum, 60, 75, 77<br />

Ramichloridium indicum, 59 t –61, 63 c , 65 c , 73 *<br />

Ramichloridium mackenziei, 58, 59 t –60 t , 75, 80, 83, 188 c , 193 c ,<br />

243 c<br />

Ramichloridium musae, 57–59 t , 60–61, 63 c , 65 c , 70 * , 72–73, 91<br />

Ramichloridium obovoideum, 60, 83<br />

Ramichloridium pini, 58–59 t , 60–61, 63 c , 65 c , 74<br />

Ramichloridium schulzeri, 58, 84<br />

Ramichloridium schulzeri var. flexuosum, 60, 84<br />

Ramichloridium schulzeri var. schulzeri, 60<br />

Ramichloridium schulzeri var. tritici, 84<br />

Ramichloridium strelitziae, 59 t , 63 c , 65 c , 73 * –74 *<br />

Ramichloridium subulatum, 60, 89<br />

Ramularia, 51 k , 164<br />

Ramularia aplospora, 34 t , 36 c –37 c<br />

Ramularia endophylla, 28<br />

Ramularia miae, 6 c –7 c , 63 c , 65 c<br />

Ramularia pratensis var. pratensis, 3 t , 6 c –7 c , 63 c , 65 c<br />

Ramularia sp., 3 t , 6 c –7 c , 63 c , 65 c<br />

Ramularia torvi, 85<br />

Pseudocladosporium, 17, 43, 52 k , 54 k –55, 185, 186, 188–189, Ramulispora, 4, 31<br />

197–198, 204–205, 216, 227, 230<br />

Pseudocladosporium brevicatenatum, 205, 212<br />

Pseudocladosporium caruanianum, 208<br />

Pseudocladosporium hachijoense, 43, 212<br />

Pseudocladosporium matsushimae, 209<br />

Pseudocladosporium proteae, 198<br />

Pseudocyphellaria perpetua, 240 c<br />

Pseudodidymaria, 87<br />

Pseudomassaria carolinensis, 37 c<br />

Pseudomicrodochium, 201<br />

Pseudomicrodochium aciculare, 201<br />

Pseudotaeniolina, 11 k –12, 24<br />

Pseudotaeniolina convolvuli, 24<br />

Pseudotaeniolina globosa, 3 t , 6 c –7 c , 24<br />

Pseudovirgaria, 61 k , 87, 92<br />

Pseudovirgaria hyperparasitica, 59 t , 63 c , 65 c , 67 * , 86 * –87<br />

Psilobotrys schulzeri, 84<br />

Pycnostysanus, 238, 244<br />

Pycnostysanus azaleae, 238<br />

Pycnostysanus resinae, 237–238 t , 239<br />

Rachicladosporium, 38–39, 52 k , 55<br />

Rachicladosporium luculiae, 34 t , 36 c –37 c , 38 * –39<br />

Racodium, 75<br />

Racodium cellare, 75<br />

Racodium resinae, 235, 237–238 t , 239<br />

Racodium resinae (β) piceum, 236<br />

Racodium rupestre, 75<br />

Radulidium, 61 k , 89, 92<br />

Radulidium epichloës, 59 t , 63 c , 65 c , 87–88 * , 89<br />

Radulidium sp., 59 t , 63 c , 65 c<br />

Radulidium subulatum, 59 t , 63 c , 65 c , 70 * , 87–88 * , 89<br />

Ramichloridium, 17, 57–58, 60–61 k , 68, 75, 84, 89, 91–92<br />

Ramichloridium anceps, 60, 75, 188 c , 193 c , 243 c<br />

Ramichloridium apiculatum, 17, 57, 59 t –61, 63 c , 65 c , 68–69 * , 73<br />

Ramichloridium australiense, 59 t , 63 c , 65 c , 69 * –70 *<br />

Ramichloridium basitonum, 58, 60, 75, 77<br />

Ramichloridium biverticillatum, 59 t , 61, 63 c , 65 c , 71 * –72, 91<br />

Ramichloridium brasilianum, 59 t , 63 c , 65 c , 70 * –71 * , 72<br />

Ramulispora sorghi, 162 c<br />

Rasutoria pseudotsugae, 63 c , 65 c<br />

Rasutoria tsugae, 63 c , 65 c<br />

Readeriella, 11 k –12 k , 17, 24–25, 30<br />

Readeriella blakelyi, 6 c –7 c , 26<br />

Readeriella brunneotingens, 3 t , 6 c –7 c , 26–27 *<br />

Readeriella considenianae, 6 c –7 c , 26, 63 c , 65 c<br />

Readeriella destructans, 3 t , 6 c –7 c , 26<br />

Readeriella dimorpha, 6 c –7 c , 26<br />

Readeriella epicoccoides, 6 c –7 c , 11, 30<br />

Readeriella eucalypti, 3 t , 6 c –7 c , 26<br />

Readeriella gauchensis, 3 t , 6 c –7 c , 26<br />

Readeriella mirabilis, 3 t , 6 c –7 c , 24–25, 27 * , 30, 97 c<br />

Readeriella molleriana, 3 t , 10<br />

Readeriella novaezel<strong>and</strong>iae, 6 c –7 c , 97 c<br />

Readeriella nubilosa, 10<br />

Readeriella ovata, 3 t , 6 c –7 c<br />

Readeriella pulcherrima, 6 c –7 c , 26<br />

Readeriella readeriellophora, 11, 25–27 *<br />

Readeriella stellenboschiana, 3 t , 6 c –7 c , 26<br />

Readeriella toledana, 11<br />

Readeriella zuluensis, 3 t , 6 c –7 c , 26<br />

Repetophragma goidanichii, 62 c , 64 c<br />

Retroconis, 46<br />

Retroconis fusiformis, 34 t , 36 c –37 c , 46<br />

Rhinocladiella, 57–58, 60–61 k , 75–76, 92<br />

Rhinocladiella anceps, 59 t , 62 c , 64 c , 75 * –76<br />

Rhinocladiella apiculata, 68<br />

Rhinocladiella atrovirens, 62 c , 64 c , 75–76, 243 c<br />

Rhinocladiella basitona, 59 t , 62 c , 64 c , 76 * –77, 243 c<br />

Rhinocladiella cellaris, 75<br />

Rhinocladiella elatior, 89<br />

Rhinocladiella fasciculata, 59 t , 62 c , 64 c , 77 *<br />

Rhinocladiella indica, 68, 73<br />

Rhinocladiella mackenziei, 62 c , 64 c , 78 * , 80, 83, 92<br />

Rhinocladiella obovoidea, 83<br />

Rhinocladiella phaeophora, 60 t , 62 c , 64 c , 76<br />

Rhinocladiella schulzeri, 84<br />

Ramichloridium cerophilum, 59 t –61, 63 c , 65 c , 71 * –72, 73, 241, Rhinocladiella similis, 243 c<br />

251


Rhinotrichum multisporum, 84<br />

Rhizocladosporium, 50–51, 54 k –55, 204<br />

Rhizocladosporium argillaceum, 34 t , 36 c –37 c , 50 * –51<br />

Rhodoveronaea, 61 k , 89–90, 92<br />

Rhodoveronaea varioseptata, 60 t , 62 c , 64 c , 70 * , 90–91<br />

Rhytisma acerinum, 240 c<br />

Saccharomyces cerevisiae, 240 c<br />

Sarcoleotia turficola, 36 c –37 c<br />

Satchmopsis brasiliensis, 189 c , 193 c<br />

Schizothyrium, 5 k , 28, 30<br />

Schizothyrium acerinum, 28<br />

Schizothyrium pomi, 6 c –7 c<br />

Scolicotrichum iridis, 125<br />

Scorias spongiosa, 6 c –7 c<br />

Seifertia, 52 k , 238, 244<br />

Seifertia azaleae, 235, 238–239, 240 c , 241–242, 244<br />

Septocylindrium chaetospira, 187<br />

Septonema, 54 k<br />

Septonema chaetospira, 187<br />

Septoria, 164, 173<br />

Septoria pulcherrima, 26<br />

Septoria tritici, 3 t , 6 c –7 c , 63 c , 65 c<br />

Setosphaeria monoceras, 240 c<br />

Sonderhenia, 24<br />

Sorocybe, 52 k , 54 k , 236–238, 242–244<br />

Sorocybe resinae, 235, 237 * –238 t , 239–240 c , 241–243 c , 244<br />

Spadicoides minuta, 187<br />

Spathularia flavida, 240 c<br />

Sphaerella allicina, 118<br />

Sphaerella cryptica, 10<br />

Sphaerella molleriana, 10<br />

Sphaerella nubilosa, 10<br />

Sphaerella tassiana, 8, 122<br />

Sphaeria allicina, 118<br />

Spilocaea, 185, 205<br />

Spilomyces dendriticus, 10<br />

Sporidesmium pachyanthicola, 63 c , 65 c<br />

Sporocybe, 236<br />

Sporocybe resinae, 237–238 t , 239<br />

Sporotrichum anceps, 76<br />

Sporotrichum nigrum, 238 t<br />

Stagonospora pulcherrima, 26<br />

Staninwardia, 12 k , 26<br />

Staninwardia breviuscula, 26<br />

Staninwardia suttonii, 5, 6 c –7 c , 26, 36 c –37 c , 63 c , 65 c<br />

Stenella, 31, 44–45, 53 k , 75, 227<br />

Stenella araguata, 16, 19, 24, 43, 44 * –45 * , 75<br />

“Stenella” cerophilum, 36 c –37 c<br />

Stenellopsis, 31<br />

Stigmidium, 30<br />

Stigmidium schaereri, 30 *<br />

Stigmina, 4, 12, 30–31<br />

Stigmina eucalypti, 12<br />

Stigmina platani, 12<br />

Strigopodia resinae, 236<br />

Stysanopsis resinae, 238 t<br />

Stysanus resinae, 237–238 t , 239<br />

Subramaniomyces, 51, 54 k , 204<br />

Subramaniomyces fusisaprophyticus, 34 t –35, 37 c , 51<br />

Subramaniomyces simplex, 51<br />

Sympodiella, 204<br />

Sympoventuria, 185, 204<br />

Sympoventuria capensis, 189 c , 191 t , 194 c , 205<br />

Taeniolella, 41, 52 k<br />

Taeniolina, 52 k<br />

Taeniolina scripta, 17<br />

Teratosphaeria, 5 k , 8, 28, 30–31, 43, 45, 92, 96<br />

Teratosphaeria africana, 8<br />

Teratosphaeria alistairii, 6 c –7 c , 9, 36 c –37 c , 63 c , 65 c<br />

Teratosphaeria associata, 9<br />

Teratosphaeria bellula, 3 t , 6 c –7 c , 10<br />

Teratosphaeria cryptica, 6 c –7 c , 10, 36 c –37 c<br />

Teratosphaeria dendritica, 2 t , 10<br />

Teratosphaeria excentrica, 10<br />

Teratosphaeria fibrillosa, 3 t –4, 6 c –7 c , 8–9 * , 10, 63 c , 65 c<br />

Teratosphaeria fimbriata, 10<br />

Teratosphaeria flexuosa, 6 c –7 c , 10<br />

Teratosphaeria gamsii, 10<br />

Teratosphaeria jonkershoekensis, 10<br />

Teratosphaeria juvenis, 36 c –37 c<br />

Teratosphaeria maxii, 6 c –7 c , 10<br />

Teratosphaeria mexicana, 3 t , 6 c –7 c , 10<br />

Teratosphaeria microspora, 2 t , 8, 10, 15 k , 97 c<br />

Teratosphaeria molleriana, 3 t , 6 c –7 c , 10, 36 c –37 c , 63 c , 65 c<br />

Teratosphaeria nubilosa, 3 t , 6 c –7 c , 10, 162 c<br />

Teratosphaeria ohnowa, 4 t , 6 c –7 c , 10<br />

Teratosphaeria parkiiaffinis, 10<br />

Teratosphaeria parva, 6 c –7 c , 10, 63 c , 65 c<br />

Teratosphaeria perpendicularis, 10<br />

Teratosphaeria pluritubularis, 10<br />

Teratosphaeria proteaarboreae, 8<br />

Teratosphaeria pseudafricana, 11<br />

Teratosphaeria pseudocryptica, 11<br />

Teratosphaeria pseudosuberosa, 2 t , 6 c –7 c , 11<br />

Teratosphaeria quasicercospora, 11<br />

Teratosphaeria readeriellophora, 6 c –7 c , 11, 63 c , 65 c , 97 c<br />

Teratosphaeria secundaria, 4 t , 6 c –7 c , 11<br />

Teratosphaeria sp., 2 t , 4 t , 6 c –7 c<br />

Teratosphaeria stramenticola, 11<br />

Teratosphaeria suberosa, 6 c –7 c , 11, 63 c , 65 c<br />

Teratosphaeria suttonii, 3 t , 6 c –7 c , 11, 36 c –37 c<br />

Teratosphaeria toledana, 6 c –7 c , 11, 63 c , 65 c<br />

Terfezia gigantea, 240 c<br />

<strong>The</strong>dgonia ligustrina, 34 t , 36 c –37 c<br />

Thielaviopsis basicola, 159 t<br />

Thyridium, 84<br />

Thyridium vestitum, 60, 62 c , 64 c<br />

Thysanorea, 61 k , 63, 80, 92<br />

Thysanorea papuana, 60 t , 62 c , 64 c , 68 * , 78 * –79 * , 80, 188 c , 193 c<br />

Torrendiella eucalypti, 36 c –37 c<br />

Torula pinophila, 239<br />

Toxicocladosporium, 39, 53 k , 55<br />

Toxicocladosporium irritans, 34 t , 36 c –37 c , 39–40 * , 41<br />

Trematosphaeria heterospora, 240 c<br />

Trichosphaeria pilosa, 6 c –7 c<br />

Trimmatostroma, 5, 14, 164<br />

Trimmatostroma abietina, 14–15, 97 c<br />

Trimmatostroma abietis, 14–15, 97 c<br />

Trimmatostroma betulinum, 3 t , 5, 6 c –7 c<br />

Trimmatostroma elginense, 16<br />

Trimmatostroma excentricum, 10<br />

Trimmatostroma macowanii, 17<br />

Trimmatostroma microsporum, 10<br />

Trimmatostroma protearum, 17<br />

252


Trimmatostroma salicis, 3 t , 5, 6 c –7 c , 13 * , 97 c<br />

Uwebraunia, 28<br />

Venturia, 43, 52 k , 55, 185–186, 204–205, 216<br />

Venturia aceris, 189 c , 191 t , 194 c<br />

Venturia alpina, 189 c , 191 t , 194 c<br />

Venturia anemones, 189 c , 191 t , 194 c<br />

Venturia asperata, 62 c , 64 c<br />

Venturia atriseda, 189 c , 191 t , 194 c<br />

Venturia aucupariae, 189 c , 191 t , 194 c<br />

Venturia carpophila, 62 c , 64 c , 189 c –190 t , 194 c<br />

Venturia cephalariae, 189 c , 191 t , 194 c<br />

Venturia cerasi, 189 c , 191 t , 194 c<br />

Venturia chlorospora, 62 c , 64 c , 189 c , 191 t , 194 c<br />

Venturia crataegi, 189 c , 191 t , 194 c<br />

Venturia dickiei, 205<br />

Venturia ditricha, 189 c , 191 t , 194 c<br />

Venturia fraxini, 189 c , 192 t , 194 c<br />

Venturia hanliniana, 62 c , 64 c , 205, 212<br />

Venturia helvetica, 189 c , 192 t , 194 c<br />

Venturia hystrioides, 189 c , 192 t , 194 c , 212–213 *<br />

Venturia inaequalis, 62 c , 64 c , 189 c , 191 t , 194 c<br />

Venturia lonicerae, 189 c , 192 t , 194 c<br />

Venturia macularis, 189 c , 192 t , 194 c<br />

Venturia maculiformis, 189 c , 192 t , 194 c<br />

Venturia m<strong>and</strong>shurica, 191 t , 210<br />

Venturia minuta, 189 c , 192 t , 194 c<br />

Venturia nashicola, 189 c , 192 t , 194 c<br />

Venturia polygonivivipari, 189 c , 192 t , 194 c<br />

Venturia populina, 189 c , 192 t , 194 c<br />

Venturia pyrina, 62 c , 64 c , 189 c , 192 t , 194 c<br />

Venturia saliciperda, 189 c , 192 t , 194 c<br />

Venturia sp., 189 c , 192 t , 194 c<br />

Venturia tremulae, 191 t<br />

Venturia tremulae var. gr<strong>and</strong>identatae, 189 c , 192 t , 194 c<br />

Venturia tremulae var. populialbae, 189 c , 192 t , 194 c<br />

Venturia tremulae var. tremulae, 189 c , 192 t , 194 c<br />

Venturia viennotii, 189 c , 192 t , 194 c<br />

Veronaea, 57–58, 60–61 k , 76, 80, 87, 89, 91, 92<br />

Veronaea apiculata, 68<br />

Veronaea botryosa, 60 t , 62 c , 64 c , 76, 80–81 * , 188 c , 193 c<br />

Veronaea compacta, 60 t , 62 c , 64 c , 81–82 * , 83<br />

Veronaea harunganae, 92<br />

Veronaea indica, 73<br />

Veronaea japonica, 60 t , 62 c , 64 c , 74 * , 82 *<br />

Veronaea musae, 57, 72<br />

Veronaea parvispora, 76<br />

Veronaea simplex, 60, 91<br />

Veronaea verrucosa, 73<br />

Veronaeopsis, 61 k , 91–92<br />

Veronaeopsis simplex, 60 t , 62 c , 64 c , 74 * , 90 * –91, 189 c , 194 c<br />

Verrucisporota, 31, 45<br />

Verrucocladosporium, 41, 53 k , 55<br />

Verrucocladosporium dirinae, 34 t , 36 c –37 c , 41–42 *<br />

Vibrissea flavovirens, 6 c –7 c<br />

Vibrissea truncorum, 6 c –7 c<br />

Virgaria, 87<br />

Websteromyces, 52 k<br />

Wentiomyces, 91<br />

Wentiomyces javanicus, 91<br />

Westerdykella cylindrica, 240 c<br />

Xenomeris juniperi, 63 c , 65 c<br />

Xylographa vitiligo, 240 c<br />

Xylohypha, 52 k<br />

Zasmidium, 45, 61, 74–75<br />

Zasmidium cellare, 60 t –61, 63 c , 65 c , 74 * –75<br />

Zeloasperisporium, 214, 216<br />

Zeloasperisporium hyphopodioides, 189 c , 191 t , 194 c , 214–215 *<br />

Zygosporium, 5 k<br />

253

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!